22.01.2015 Views

hf-report20121214

hf-report20121214

hf-report20121214

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

EPA 601/R-12/011 | December 2012 | www.epa.gov/<strong>hf</strong>study<br />

Study of the Potential Impacts of<br />

Hydraulic Fracturing on<br />

Drinking Water Resources<br />

PROGRESS REPORT<br />

United States Environmental Protection Agency<br />

Office of Research and Development


this page intentially left blank


Study of the Potential Impacts of<br />

Hydraulic Fracturing on<br />

Drinking Water Resources<br />

PROGRESS REPORT<br />

US Environmental Protection Agency<br />

Office of Research and Development<br />

Washington, DC<br />

December 2012<br />

EPA/601/R-12/011


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Disclaimer<br />

Mention of trade names or commercial products does not constitute<br />

endorsement or recommendation for use.<br />

ii


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table of Contents<br />

Executive Summary .................................................................................................................................... 1 <br />

1. Introduction ......................................................................................................................................... 5 <br />

1.1. Stakeholder Engagement ...................................................................................................... 6 <br />

2. Overview of the Research Program .................................................................................................. 8 <br />

2.1. Research Questions ............................................................................................................ 12<br />

2.2. Environmental Justice ......................................................................................................... 21<br />

2.3. Changes to the Research Program..................................................................................... 22<br />

2.4. Research Approach ............................................................................................................ 23<br />

3. Analysis of Existing Data ................................................................................................................. 25<br />

3.1. Literature Review ................................................................................................................ 25<br />

3.2. Spills Database Analysis ..................................................................................................... 31<br />

3.3. Service Company Analysis ................................................................................................. 39<br />

3.4. Well File Review .................................................................................................................. 46<br />

3.5. FracFocus Analysis ............................................................................................................. 54<br />

4. Scenario Evaluations ........................................................................................................................ 62<br />

4.1. Subsurface Migration Modeling ........................................................................................... 62<br />

4.2. Surface Water Modeling ...................................................................................................... 75<br />

4.3. Water Availability Modeling ................................................................................................. 80<br />

5. Laboratory Studies ........................................................................................................................... 94<br />

5.1. Source Apportionment Studies ........................................................................................... 94<br />

5.2. Wastewater Treatability Studies ........................................................................................ 101<br />

5.3. Brominated Disinfection Byproduct Precursor Studies ..................................................... 107<br />

5.4. Analytical Method Development ........................................................................................ 112<br />

6. Toxicity Assessment ...................................................................................................................... 122<br />

7. Case Studies .................................................................................................................................... 127<br />

7.1. Introduction to Case Studies ............................................................................................. 127<br />

7.2. Las Animas and Huerfano Counties, Colorado ................................................................. 131<br />

7.3. Dunn County, North Dakota .............................................................................................. 137<br />

7.4. Bradford County, Pennsylvania ......................................................................................... 142<br />

7.5. Washington County, Pennsylvania ................................................................................... 148<br />

7.6. Wise County, Texas .......................................................................................................... 153<br />

8. Conducting High-Quality Science ................................................................................................. 159<br />

8.1. Quality Assurance ............................................................................................................. 159<br />

8.2. Peer Review ...................................................................................................................... 161<br />

9. Research Progress Summary and Next Steps ............................................................................. 163<br />

9.1. Summary of Progress by Research Activity ...................................................................... 163<br />

9.2. Summary of Progress by Water Cycle Stage ................................................................... 165<br />

iii


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table of Contents<br />

9.3. Report of Results............................................................................................................... 170<br />

9.4. Conclusions ....................................................................................................................... 170<br />

10. References ....................................................................................................................................... 172<br />

Appendix A: Chemicals Identified in Hydraulic Fracturing Fluids and Wastewater ........................ 196<br />

Appendix B: Stakeholder Engagement ................................................................................................. 246<br />

Appendix C: Summary of QAPPs .......................................................................................................... 251<br />

Appendix D: Divisions of Geologic Time .............................................................................................. 253<br />

Glossary ................................................................................................................................................... 254<br />

iv


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

List of Tables <br />

Table 1. Titles and descriptions of the research projects conducted as part of the EPA’s Study of the<br />

Potential Impacts of Hydraulic Fracturing on Drinking Water Resources ................................................... 10<br />

Table 2. Secondary research questions and applicable research projects identified for the water<br />

acquisition stage of the hydraulic fracturing water cycle ............................................................................. 15<br />

Table 3. Secondary research questions and applicable research projects identified for the chemical<br />

mixing stage of the hydraulic fracturing water cycle ................................................................................... 16<br />

Table 4. Secondary research questions and applicable research projects identified for the well injection<br />

stage of the hydraulic fracturing water cycle. .............................................................................................. 17<br />

Table 5. Secondary research questions and applicable research projects identified for the flowback and<br />

produced water stage of the hydraulic fracturing water cycle ..................................................................... 19<br />

Table 6. Secondary research questions and applicable research projects identified for the wastewater<br />

treatment and waste disposal stage of the hydraulic fracturing water cycle ............................................... 21<br />

Table 7. Research questions addressed by assessing the demographics of locations where hydraulic<br />

fracturing activities are underway ............................................................................................................... 21<br />

Table 8. Research activities and objectives ............................................................................................... 24<br />

Table 9. Classifications of information sources with examples .................................................................. 26<br />

Table 10. Description of factors used to assess the quality of existing data and information compiled <br />

during the literature review .......................................................................................................................... 27<br />

Table 11. Chemicals identified by the US House of Representatives Committee on Energy and <br />

Commerce as known or suspected carcinogens, regulated under the Safe Drinking Water Act (SDWA) or<br />

classified as hazardous air pollutants (HAP) under the Clean Air Act ........................................................ 29<br />

Table 12. Chemical appearing most often in hydraulic fracturing in over 2,500 products reported by 14<br />

hydraulic fracturing service companies as being used between 2005 and 2009 ....................................... 29<br />

Table 13. Secondary research questions addressed by reviewing existing databases that contain data<br />

relating to surface spills of hydraulic fracturing fluids and wastewater ....................................................... 31<br />

Table 14. Oil and gas-related spill databases used to compile information on hydraulic fracturing-related <br />

incidents ...................................................................................................................................................... 32<br />

Table 15. Data fields available in the NRC Freedom of Information Act database .................................... 34<br />

Table 16. Preset search terms available for the spill material, spill cause, and spill source data fields in<br />

the New Mexico Oil Conservation Division Spills Database ....................................................................... 36<br />

Table 17. Total number of incidents retrieved from the Pennsylvania Department of Environmental<br />

Protection's Compliance Reporting Database by varying inputs in the “Marcellus only” and inspections<br />

with “violations only data fields.” ................................................................................................................. 37<br />

Table 18. Secondary research questions addressed by analyzing data received from nine hydraulic<br />

fracturing service companies ...................................................................................................................... 39<br />

Table 19. Annual revenue and approximate number of employees for the nine service companies<br />

selected to receive the EPA’s September 2010 information request ......................................................... 40<br />

Table 20. Formulations, products, and chemicals reported as used or distributed by the nine service<br />

companies between September 2005 and September 2010 ...................................................................... 45<br />

Table 21. Secondary research questions addressed by the well file review research project ................... 46<br />

Table 22. The potential relationship between the topic areas in the information request and the stages of<br />

the hydraulic fracturing water cycle ............................................................................................................. 50<br />

v


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

List of Tables<br />

Table 23. Number of wells for which data were provided by each operator .............................................. 51<br />

Table 24. Secondary research questions addressed by extracting data from FracFocus, a nationwide <br />

hydraulic fracturing chemical registry .......................................................................................................... 54<br />

Table 25. Number of wells, by state, with data in FracFocus as of February 2012 ................................... 60<br />

Table 26. Secondary research questions addressed by simulating the subsurface migration of gases and <br />

fluids resulting from six possible mechanisms ............................................................................................ 62<br />

Table 27. Modules combined with the Transport of Unsaturated Groundwater and Heat (TOUGH)......... 71<br />

Table 28. Secondary research question addressed by modeling surface water discharges from<br />

wastewater treatment facilities accepting hydraulic fracturing wastewater ................................................ 75<br />

Table 29. Research questions addressed by modeling water withdrawals and availability in selected river<br />

basins .......................................................................................................................................................... 80<br />

Table 30. Water withdrawals for use in the Susquehanna River Basin ..................................................... 83<br />

Table 31. Well completions for select counties in Colorado within the Upper Colorado River Basin <br />

watershed .................................................................................................................................................... 86<br />

Table 32. Water withdrawals for use in the Upper Colorado River Basin .................................................. 86<br />

Table 33. Estimated total annual water demand for oil and gas wells in Colorado that were hydraulically <br />

fractured in 2010 and 2011 ......................................................................................................................... 88<br />

Table 34. Data and assumptions for future watershed availability and use scenarios modeled for the <br />

Susquehanna River Basin........................................................................................................................... 91<br />

Table 35. Data and assumptions for future watershed availability and use scenarios modeled for the <br />

Upper Colorado River Basin ....................................................................................................................... 91<br />

Table 36. Secondary research questions addressed by the source apportionment research project ....... 94<br />

Table 37. Historical average of monthly mean river flow and range of monthly means from 2006 <br />

through 2011 for two rivers in Pennsylvania where the EPA collects samples for source apportionment<br />

research ...................................................................................................................................................... 96<br />

Table 38. Distance between sampling sites and wastewater treatment facilities on two rivers where the <br />

EPA collects samples for source apportionment research ......................................................................... 96<br />

Table 39. Inorganic analyses and respective instrumentation planned for source apportionment<br />

research ...................................................................................................................................................... 97<br />

Table 40. Median concentrations of selected chemicals and conductivity of effluent treated and <br />

discharged from two wastewater treatment facilities that accept oil and gas wastewater .......................... 99<br />

Table 41. Secondary research questions addressed by the wastewater treatability laboratory<br />

studies ....................................................................................................................................................... 101<br />

Table 42. Chemicals identified for initial studies on the adequacy of treatment of hydraulic fracturing <br />

wastewaters by conventional publicly owned treatment works, commercial treatment systems, and water<br />

reuse systems ........................................................................................................................................... 106<br />

Table 43. Secondary research questions potentially answered by studying brominated DBP formation <br />

from treated hydraulic fracturing wastewater ............................................................................................ 107<br />

Table 44. Disinfection byproducts regulated by the National Primary Drinking Water Regulations. ........ 109<br />

Table 45. Chemicals identified for analytical method testing activities .................................................... 114<br />

Table 46. Existing standard methods for analysis of selected hydraulic fracturing-related chemicals listed<br />

in Table 45................................................................................................................................................. 117<br />

vi


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

List of Tables <br />

Table 47. Secondary research questions addressed by compiling existing information on hydraulic<br />

fracturing-related chemicals ...................................................................................................................... 122<br />

Table 48. References used to develop a consolidated list of chemicals reportedly used in hydraulic<br />

fracturing fluids and/or found in flowback and produced water................................................................. 123<br />

Table 49. Secondary research questions addressed by conducting case studies................................... 127<br />

Table 50. General approach for conducting retrospective case studies .................................................. 129<br />

Table 51. Analyte groupings and examples of chemicals measured in water samples collected at the <br />

retrospective case study locations ............................................................................................................ 130<br />

Table 52. Background water quality data for the Killdeer Aquifer in North Dakota .................................. 139<br />

Table 53. Background (pre-drill) water quality data for ground water wells in Bradford County,<br />

Pennsylvania ............................................................................................................................................. 144<br />

Table 54. Background water quality data for all of Wise County, Texas, and its northern and southern<br />

regions ....................................................................................................................................................... 155<br />

Table A-1. List of CASRNs and names of chemicals reportedly used in hydraulic fracturing fluids ........ 197<br />

Table A-2. List of generic names of chemicals reportedly used in hydraulic fracturing fluids .................. 229<br />

Table A-3. List of CASRNs and names of chemicals detected in hydraulic fracturing wastewater ......... 240<br />

Table A-4. List of chemicals and properties detected in hydraulic fracturing wastewater ....................... 244<br />

Table C-1. QAPPs associated with the research projects discussed in this progress report. ................. 251<br />

vii


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

List of Figures <br />

Figure 1. Illustration of the five stages of the hydraulic fracturing water cycle ............................................. 8 <br />

Figure 2. Potential drinking water issues associated with each stage of the hydraulic fracturing water<br />

cycle .............................................................................................................................................................. 9 <br />

Figure 3. Illustration of the structure of the EPA’s Study of the Potential Impacts of Hydraulic Fracturing <br />

on Drinking Water Resources ..................................................................................................................... 12<br />

Figure 4. Fundamental research questions posed for each stage of the hydraulic fracturing water<br />

cycle ............................................................................................................................................................ 13<br />

Figure 5. Water acquisition......................................................................................................................... 14<br />

Figure 6. Chemical mixing .......................................................................................................................... 15<br />

Figure 7. Well injection ............................................................................................................................... 17<br />

Figure 8. Flowback and produced water .................................................................................................... 18<br />

Figure 9. Wastewater treatment and waste disposal ................................................................................. 20<br />

Figure 10. Locations of oil and gas production wells hydraulically fractured between September 2009 <br />

and October 2010 ....................................................................................................................................... 44<br />

Figure 11. Locations of oil and gas production wells hydraulically fractured from September 2009 <br />

through October 2010 ................................................................................................................................. 49<br />

Figure 12. Locations of 333 wells selected for the well file review............................................................. 52<br />

Figure 13. Example of data disclosed through FracFocus ......................................................................... 57<br />

Figure 14. Scenario A of the subsurface migration modeling project ........................................................ 64<br />

Figure 15. Scenario B1 of the subsurface migration modeling project ...................................................... 65<br />

Figure 16. Scenario B2 of the subsurface migration modeling project ...................................................... 66<br />

Figure 17. Scenario C of the subsurface migration modeling project ........................................................ 67<br />

Figure 18. Scenario D1 of the subsurface migration modeling project ...................................................... 68<br />

Figure 19. Scenario D2 of the subsurface migration modeling project ...................................................... 69<br />

Figure 20. The Susquehanna River Basin, overlying a portion of the Marcellus Shale, is one of two study <br />

areas chosen for water availability modeling .............................................................................................. 81<br />

Figure 21. The Upper Colorado River Basin, overlying a portion of the Piceance Basin, is one of two river<br />

basins chosen for water availability modeling ............................................................................................. 82<br />

Figure 22. Public water systems in the Susquehanna River Basin............................................................ 84<br />

Figure 23. Public water systems in the Upper Colorado River Basin ........................................................ 87<br />

Figure 24. Hydraulic fracturing wastewater flow in unconventional oil and gas extraction ...................... 102<br />

Figure 25. Generalized flow diagram for conventional publicly owned works treatment processes ........ 103<br />

Figure 26. Flow diagram of the EPA’s process leading to the development of modified or new analytical <br />

methods..................................................................................................................................................... 116<br />

Figure 27. Locations of the five retrospective case studies chosen for inclusion in the EPA’s Study of the <br />

Potential Impacts of Hydraulic Fracturing on Drinking Water Resources ................................................. 128<br />

Figure 28. Extent of the Raton Basin in southeastern Colorado and northeastern New Mexico ............. 132<br />

Figure 29. Locations of sampling sites in Las Animas and Huerfano Counties, Colorado ...................... 135<br />

Figure 30. Extent of the Bakken Shale in North Dakota and Montana .................................................... 137<br />

viii


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

List of Figures<br />

Figure 31. Location of sampling sites in Dunn County, North Dakota ..................................................... 140<br />

Figure 32. Extent of the Marcellus Shale, which underlies large portions of New York, Ohio,<br />

Pennsylvania, and West Virginia .............................................................................................................. 142<br />

Figure 33. Location of sampling sites in Bradford and Susquehanna Counties, Pennsylvania ............... 146<br />

Figure 34. Extent of the Marcellus Shale, which underlies large portions of New York, Ohio,<br />

Pennsylvania, and West Virginia .............................................................................................................. 148<br />

Figure 35. Sampling locations in Washington County, Pennsylvania ...................................................... 151<br />

Figure 36. Extent of the Barnett Shale in north-central Texas ................................................................. 153<br />

Figure 37. Location of sampling sites in Wise County, Texas ................................................................. 157<br />

Figure 38a. Summary of research projects underway for the first three stages of the hydraulic fracturing <br />

water cycle ................................................................................................................................................ 166<br />

Figure 38b. Summary of research projects underway for the first three stages of the hydraulic fracturing <br />

water cycle ................................................................................................................................................ 167<br />

Figure 39a. Summary of research projects underway for the last two stages of the hydraulic fracturing <br />

water cycle ................................................................................................................................................ 168<br />

Figure 39b. Summary of research projects underway for the last two stages of the hydraulic fracturing <br />

water cycle ................................................................................................................................................ 169<br />

Figure B-1. Timeline for technical roundtables and workshops ............................................................... 250<br />

Figure D-1. Divisions of geologic time approved by the USGS Geologic Names Committee (2010) ...... 253<br />

ix


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

ADQ<br />

API<br />

ASTM<br />

Br-DBP<br />

BTEX<br />

CASRN<br />

CBI<br />

CBM<br />

COGCC<br />

CWT<br />

DBP<br />

DSSTox<br />

FORTRAN<br />

GIS<br />

GWPC<br />

HAA<br />

HSPF<br />

IRIS<br />

LBNL<br />

LOAEL<br />

MCL<br />

MGD<br />

MSDS<br />

NAS<br />

NDIC<br />

NEMS<br />

NOM<br />

NPDES<br />

NRC<br />

NYSDEC<br />

PADEP<br />

POTW<br />

PPRTV<br />

PWS<br />

QA<br />

List of Acronyms and Abbreviations<br />

Audit of data quality<br />

American Petroleum Institute<br />

American Society for Testing and Materials<br />

Brominated disinfection byproduct<br />

Benzene, toluene, ethylbenzene, and xylene<br />

Chemical Abstracts Service Registration Number<br />

Confidential business information<br />

Coalbed methane<br />

Colorado Oil and Gas Conservation Commission<br />

Centralized waste treatment facility<br />

Disinfection byproduct<br />

Distributed Structure-Searchable Toxicity Database Network<br />

Formula translation<br />

Geographic information system<br />

Ground Water Protection Council<br />

Haloacetic acid<br />

Hydrologic Simulation Program FORTRAN<br />

Integrated Risk Information System<br />

Lawrence Berkeley National Laboratory<br />

Lowest observed adverse effect levels<br />

Maximum contaminant level<br />

Million gallons per day<br />

Material Safety Data Sheet<br />

National Academy of Sciences<br />

North Dakota Industrial Commission<br />

National Energy Modeling System<br />

Naturally occurring organic matter<br />

National Pollutant Discharge Elimination System<br />

National Response Center<br />

New York State Department of Environmental Conservation<br />

Pennsylvania Department of Environmental Protection<br />

Publicly owned treatment work<br />

Provisional Peer-Reviewed Toxicity Value<br />

Public water systems<br />

Quality assurance<br />

x


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

QAPP<br />

QC<br />

RRC<br />

SDWA<br />

SOP<br />

SRB<br />

SRBC<br />

SWAT<br />

TDS<br />

THM<br />

TOPKAT<br />

TOUGH<br />

TSA<br />

TSCA<br />

UCRB<br />

UIC<br />

US EIA<br />

US EPA<br />

US FWS<br />

US GAO<br />

US OMB<br />

USCB<br />

USDA<br />

USGS<br />

USHR<br />

WWTF<br />

List of Acronyms and Abbreviations<br />

Quality assurance project plan<br />

Quality control<br />

Railroad Commission of Texas<br />

Safe Drinking Water Act<br />

Standard operating procedure<br />

Susquehanna River Basin<br />

Susquehanna River Basin Commission<br />

Soil and Water Assessment Tool<br />

Total dissolved solids<br />

Trihalomethane<br />

Toxicity Prediction by Komputer Assisted Technology<br />

Transport of Unsaturated Groundwater and Heat<br />

Technical systems audit<br />

Toxic Substances Control Act<br />

Upper Colorado River Basin<br />

Underground injection control<br />

US Energy Information Administration<br />

US Environmental Protection Agency<br />

US Fish and Wildlife Service<br />

US Government Accountability Office<br />

US Office of Management and Budget<br />

US Census Bureau<br />

US Department of Agriculture<br />

US Geological Survey<br />

US House of Representatives<br />

Wastewater treatment facility<br />

xi


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Executive Summary<br />

Natural gas plays a key role in our nation’s clean energy future. The United States has vast reserves<br />

of natural gas that are commercially viable as a result of advances in horizontal drilling and<br />

hydraulic fracturing technologies, which enable greater access to gas in rock formations deep<br />

underground. These advances have spurred a significant increase in the production of both natural<br />

gas and oil across the country.<br />

Responsible development of America’s oil and gas resources offers important economic, energy<br />

security, and environmental benefits. However, as the use of hydraulic fracturing has increased, so<br />

have concerns about its potential human health and environmental impacts, especially for drinking<br />

water. In response to public concern, the US House of Representatives requested that the US<br />

Environmental Protection Agency (EPA) conduct scientific research to examine the relationship<br />

between hydraulic fracturing and drinking water resources (USHR, 2009).<br />

In 2011, the EPA began research under its Plan to Study the Potential Impacts of Hydraulic<br />

Fracturing on Drinking Water Resources. The purpose of the study is to assess the potential impacts<br />

of hydraulic fracturing on drinking water resources, if any, and to identify the driving factors that<br />

may affect the severity and frequency of such impacts. Scientists are focusing primarily on<br />

hydraulic fracturing of shale formations to extract natural gas, with some study of other oil- and<br />

gas-producing formations, including tight sands, and coalbeds. The EPA has designed the scope of<br />

the research around five stages of the hydraulic fracturing water cycle. Each stage of the cycle is<br />

associated with a primary research question:<br />

• Water acquisition: What are the possible impacts of large volume water withdrawals from<br />

ground and surface waters on drinking water resources<br />

• Chemical mixing: What are the possible impacts of hydraulic fracturing fluid surface spills<br />

on or near well pads on drinking water resources<br />

• Well injection: What are the possible impacts of the injection and fracturing process on<br />

drinking water resources<br />

• Flowback and produced water: What are the possible impacts of flowback and produced<br />

water (collectively referred to as “hydraulic fracturing wastewater”) surface spills on or<br />

near well pads on drinking water resources<br />

• Wastewater treatment and waste disposal: What are the possible impacts of inadequate<br />

treatment of hydraulic fracturing wastewater on drinking water resources<br />

This report describes 18 research projects underway to answer these research questions and<br />

presents the progress made as of September 2012 for each of the projects. Information presented<br />

as part of this report cannot be used to draw conclusions about potential impacts to drinking water<br />

resources from hydraulic fracturing. The research projects are organized according to five different<br />

types of research activities: analysis of existing data, scenario evaluations, laboratory studies,<br />

toxicity assessments, and case studies.<br />

1


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Analysis of Existing Data<br />

Data from multiple sources have been obtained for review and analysis. Many of the data come<br />

directly from the oil and gas industry and states with high levels of oil and gas activity. Information<br />

on the chemicals and practices used in hydraulic fracturing has been collected from nine companies<br />

that hydraulically fractured a total of 24,925 wells between September 2009 and October 2010.<br />

Additional data on chemicals and water use for hydraulic fracturing are being pulled from over<br />

12,000 well-specific chemical disclosures in FracFocus, a national hydraulic fracturing chemical<br />

registry operated by the Ground Water Protection Council and the Interstate Oil and Gas Compact<br />

Commission. Well construction and hydraulic fracturing records provided by well operators are<br />

being reviewed for 333 oil and gas wells across the United States; data within these records are<br />

being scrutinized to assess the effectiveness of current well construction practices at containing<br />

gases and liquids before, during, and after hydraulic fracturing.<br />

Data on causes and volumes of spills of hydraulic fracturing fluids and wastewater are being<br />

collected and reviewed from state spill databases in Colorado, New Mexico, and Pennsylvania.<br />

Similar information is being collected from the National Response Center national database of oil<br />

and chemical spills.<br />

In addition, the EPA is reviewing scientific literature relevant to the research questions posed in<br />

this study. A Federal Register notice was published on November 9, 2012, requesting relevant, peerreviewed<br />

data and published reports, including information on advances in industry practices and<br />

technologies. This body of literature will be synthesized with results from the other research<br />

projects to create a report of results.<br />

Scenario Evaluations<br />

Computer models are being used to identify conditions that may lead to impacts on drinking water<br />

resources from hydraulic fracturing. The EPA has identified hypothetical, but realistic, scenarios<br />

pertaining to the water acquisition, well injection, and wastewater treatment and waste disposal<br />

stages of the water cycle. Potential impacts to drinking water sources from withdrawing large<br />

volumes of water in semi-arid and humid river basins—the Upper Colorado River Basin in the west<br />

and the Susquehanna River Basin in the east—are being compared and assessed.<br />

Additionally, complex computer models are being used to explore the possibility of subsurface gas<br />

and fluid migration from deep shale formations to overlying aquifers in six different scenarios.<br />

These scenarios include poor well construction and hydraulic communication via fractures (natural<br />

and created) and nearby existing wells. As a first step, the subsurface migration simulations will<br />

examine realistic scenarios to assess the conditions necessary for hydraulic communication rather<br />

than the probability of migration occurring.<br />

In a separate research project, concentrations of bromide and radium at public water supply<br />

intakes located downstream from wastewater treatment facilities discharging treated hydraulic<br />

fracturing wastewater are being estimated using surface water transport models.<br />

2


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Laboratory Studies<br />

Laboratory studies are largely focused on identifying potential impacts of inadequately treating<br />

hydraulic fracturing wastewater and discharging it to rivers. Experiments are being designed to test<br />

how well common wastewater treatment processes remove selected contaminants from hydraulic<br />

fracturing wastewater, including radium and other metals. Other experiments are assessing<br />

whether or not hydraulic fracturing wastewater may contribute to the formation of disinfection<br />

byproducts during common drinking water treatment processes, with particular focus on the<br />

formation of brominated disinfection byproducts, which have significant health concerns at high<br />

exposure levels.<br />

Samples of raw hydraulic fracturing wastewater, treated wastewater, and water from rivers<br />

receiving treated hydraulic fracturing wastewater have been collected for source apportionment<br />

studies. Results from laboratory analyses of these samples are being used to develop a method for<br />

determining if treated hydraulic fracturing wastewater is contributing to high chloride and bromide<br />

levels at downstream public water supplies.<br />

Finally, existing analytical methods for selected chemicals are being tested, modified, and verified<br />

for use in this study and by others, as needed. Methods are being modified in cases where standard<br />

methods do not exist for the low-level detection of chemicals of interest or for use in the complex<br />

matrices associated with hydraulic fracturing wastewater. Analytical methods are currently being<br />

tested and modified for several classes of chemicals, including glycols, acrylamides, ethoxylated<br />

alcohols, disinfection byproducts, radionuclides, and inorganic chemicals.<br />

Toxicity Assessments<br />

The EPA has identified chemicals reportedly used in hydraulic fracturing fluids from 2005 to 2011<br />

and chemicals found in flowback and produced water. Appendix A contains tables with over 1,000<br />

of these chemicals identified. Chemical, physical, and toxicological properties are being compiled<br />

for chemicals with known chemical structures. Existing models are being used to estimate<br />

properties in cases where information is lacking. At this time, the EPA has not made any judgment<br />

about the extent of exposure to these chemicals when used in hydraulic fracturing fluids or found in<br />

hydraulic fracturing wastewater, or their potential impacts on drinking water resources.<br />

Case Studies<br />

Two rounds of sampling at five case study locations in Colorado, North Dakota, Pennsylvania, and<br />

Texas have been completed. In total, water samples have been collected from over 70 domestic<br />

water wells, 15 monitoring wells, and 13 surface water sources, among others. This research will<br />

help to identify the source of any contamination that may have occurred.<br />

The EPA continues to work with industry partners to begin research activities at potential<br />

prospective case study locations, which involve sites where the research will begin before well<br />

construction. This will allow the EPA to collect baseline water quality data in the area. Water quality<br />

will be monitored for any changes throughout drilling, injection of fracturing fluids, flowback, and<br />

production. Samples of flowback and produced water will be used for other parts of the study, such<br />

as assessing the efficacy of wastewater treatment processes at removing contaminants in hydraulic<br />

fracturing wastewater.<br />

3


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Invigorating the Research Study Through Consultation and Peer Review<br />

The EPA is committed to conducting a study that uses the best available science, independent<br />

sources of information, and a transparent, peer-reviewed process that will ensure the validity and<br />

accuracy of the results. The agency is working in consultation with other federal agencies, state and<br />

interstate regulatory agencies, industry, non-governmental organizations, and others in the private<br />

and public sector. In addition to workshops held in 2011, stakeholders and technical experts are<br />

being engaged through technical roundtables and workshops, with the first set of roundtables held<br />

November 14–16, 2012. These activities will provide the EPA with ongoing access to a broad range<br />

of expertise and data, timely and constructive technical feedback, and updates on changes in<br />

industry practices and technologies relevant to the study. Technical roundtables and workshops<br />

will be followed by webinars for the general public and posting of summaries on the study’s<br />

website. Increased stakeholder engagement will also allow the EPA to educate and inform the<br />

public of the study’s goals, design, and progress.<br />

To ensure scientifically defensible results, each research project is subjected to quality assurance<br />

and peer review activities. Specific quality assurance activities performed by the EPA make sure<br />

that the agency’s environmental data are of sufficient quantity and quality to support the data’s<br />

intended use. Research products, such as papers or reports, will be subjected to both internal and<br />

external peer review before publication, which make certain that the data are used appropriately.<br />

Published results from the research projects will be synthesized in a report of results that will<br />

inform the research questions associated with each stage of the hydraulic fracturing water cycle.<br />

The EPA has designated the report of results as a “Highly Influential Scientific Assessment,” which<br />

will undergo peer review by the EPA’s Science Advisory Board, an independent and external federal<br />

advisory committee that conducts peer reviews of significant EPA research products and activities.<br />

The EPA will seek input from individual members of an ad hoc expert panel convened under the<br />

auspices of the EPA Science Advisory Board. The EPA will consider feedback from the individual<br />

experts in the development of the report of results.<br />

Ultimately, the results of this study are expected to inform the public and provide decision-makers<br />

at all levels with high-quality scientific knowledge that can be used in decision-making processes.<br />

Looking Forward: From This Report to the Next<br />

Progress<br />

Report<br />

Science<br />

Advisory<br />

Board<br />

Individual<br />

Reports<br />

and Papers<br />

Draft<br />

Report of<br />

Results<br />

Science<br />

Advisory<br />

Board Peer<br />

Review<br />

Final<br />

Report of<br />

Results<br />

Technical Roundtables and Workshops,<br />

Public Webinars<br />

4


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

1. Introduction<br />

Oil and natural gas provided more energy in the United States for residential and industrial use<br />

than any other energy source in 2010—37% and 25%, respectively (US EIA, 2011a). Advances in<br />

technology and new applications of existing techniques, as well as supportive domestic energy<br />

policy and economic developments, have recently spurred an increase in oil and gas production<br />

across a wide range of geographic regions and geologic formations in the United States. Hydraulic<br />

fracturing is a technique used to produce economically viable quantities of oil and natural gas,<br />

especially from unconventional reservoirs, such as shale, tight sands, coalbeds, and other<br />

formations. Hydraulic fracturing involves the injection of fluids under pressures great enough to<br />

fracture the oil- and gas-producing formations. The resulting fractures are held open using<br />

“proppants,” such as fine grains of sand or ceramic beads, to allow oil and gas to flow from small<br />

pores within the rock to the production well.<br />

As the use of hydraulic fracturing has increased, so have concerns about its potential impact on<br />

human health and the environment, especially with regard to possible impacts on drinking water<br />

resources. 1 These concerns have increased as oil and gas exploration and development has spread<br />

from areas with a long history of conventional production to new areas with unconventional<br />

reservoirs, such as the Marcellus Shale, which extends from New York through parts of<br />

Pennsylvania, West Virginia, eastern Ohio, and western Maryland.<br />

In response to public concerns and anticipated growth in the oil and gas industries, the US Congress<br />

urged the US Environmental Protection Agency (EPA) to examine the relationship between<br />

hydraulic fracturing and drinking water resources (USHR, 2009):<br />

The conferees urge the agency to carry out a study on the relationship between hydraulic<br />

fracturing and drinking water, using a credible approach that relies on the best available<br />

science, as well as independent sources of information. The conferees expect the study to be<br />

conducted through a transparent, peer-reviewed process that will ensure the validity and<br />

accuracy of the data. The Agency shall consult with other federal agencies as well as<br />

appropriate state and interstate regulatory agencies in carrying out the study, which should<br />

be prepared in accordance with the agency’s quality assurance principles.<br />

In 2010, the EPA launched the planning of the current study and included multiple opportunities<br />

for the public and the Science Advisory Board 2 to provide input during the study planning process. 3<br />

The EPA’s Plan to Study the Potential Impacts of Hydraulic Fracturing on Drinking Water Resources<br />

1<br />

Common concerns raised by stakeholders include potential impacts to air quality and ecosystems as well as sociologic<br />

effects (e.g., community changes). A more comprehensive list of concerns reported to the EPA during initial stakeholder<br />

meetings can be found in Appendix C of the EPA’s Plan to Study the Potential Impacts of Hydraulic Fracturing on Drinking<br />

Water Resources (EPA/600/R-11/121).<br />

2<br />

The Science Advisory Board is an independent and external federal advisory committee that conducts peer reviews of<br />

scientific matters for the EPA.<br />

3<br />

During summer 2010, the EPA engaged stakeholders in a dialogue about the study through facilitated meetings.<br />

Summaries of these meetings are available at http://www.epa.gov/<strong>hf</strong>study/publicoutreach.html.<br />

5


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

(subsequently referred to as the “Study Plan”) was finalized in November 2011 (US EPA, 2011e).<br />

The purpose of the EPA’s current study is to assess the potential impacts of hydraulic fracturing on<br />

drinking water resources, 4 if any, and to identify the driving factors that may affect the severity and<br />

frequency of such impacts. This study includes research on hydraulic fracturing to extract oil and<br />

gas from shale, tight sand, and coalbeds, focusing primarily on hydraulic fracturing of shale for gas<br />

extraction. It is intended to assess the potential impacts to drinking water resources from hydraulic<br />

fracturing as it is currently practiced and has been practiced in the past, and it is not intended to<br />

evaluate best management practices or new technologies. Emphasis is placed on identifying<br />

possible exposure pathways and hazards, providing results that can then be used to assess the<br />

potential risks to drinking water resources from hydraulic fracturing. Ultimately, results from the<br />

study are intended to inform the public and provide policymakers at all levels with high-quality<br />

scientific knowledge that can be used in decision-making.<br />

The body of this progress report presents the research progress made by the EPA, as of September<br />

2012, regarding the potential impacts of hydraulic fracturing on drinking water resources;<br />

information presented as part of this report cannot be used to draw conclusions about the<br />

proposed research questions. Chapters 3 through 7 provide project-specific updates that include<br />

background information on the research project, a description of the research methods, an update<br />

on the current status and next steps of the work, as well as a summary of the quality assurance (QA)<br />

activities to date; 5 these chapters are written for scientific and engineering professionals. All<br />

projects described in this progress report are currently underway, and nearly all are expected to be<br />

completed in the next few years. Results from individual projects will undergo peer review prior to<br />

publication. The EPA intends to synthesize the published results from these research projects in a<br />

report of results, described in more detail in Section 9.3.<br />

1.1. Stakeholder Engagement<br />

The EPA is committed to conducting this study in an open and transparent manner. During the<br />

development of the study, the EPA met with stakeholders from the general public; federal, state,<br />

regional and local agencies; tribes; industry; academia; and non-governmental organizations.<br />

Webinars and meetings with these separate groups were held to discuss the study scope, data gaps,<br />

opportunities for sharing data and conducting joint studies, current policies and practices for<br />

protecting drinking water resources, and the public engagement process.<br />

In addition to webinars and meetings, the EPA held a series of technical workshops in early 2011 on<br />

four subjects integral to hydraulic fracturing and the study: chemical and analytical methods, well<br />

construction and operation, chemical fate and transport, and water resource management. 6<br />

Technical experts from the oil and natural gas industry, academia, consulting firms, commercial<br />

laboratories, state and federal agencies, and environmental organizations were chosen to<br />

4 For this study, “drinking water resources” are considered to be any body of water, ground or surface, that could (now or<br />

in the future) serve as a source of drinking water for public or private water supplies.<br />

5 QA activities include implementation of quality assurance project plans (QAPPs), technical systems audits (TSAs), and<br />

audits of data quality (ADQs). These activities are described further in Section 8.1.<br />

6 Proceedings from the four technical workshops are available at http://www.epa.gov/<strong>hf</strong>study/technicalworkshops.html.<br />

6


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

participate in each of the workshops. The workshops gave EPA scientists the opportunity to<br />

interact with technical experts regarding current hydraulic fracturing technology and practices and<br />

to identify and design research related to the potential impacts of hydraulic fracturing on drinking<br />

water resources. Information presented during the workshops is being used to inform ongoing<br />

research.<br />

The EPA has recently announced additional opportunities for stakeholder engagement. The goals of<br />

this enhanced engagement process are to improve public understanding of the study, ensure that<br />

the EPA is current on changes in industry practices and technologies so that the report of results<br />

reflects an up-to-date picture of hydraulic fracturing operations, and obtain timely and constructive<br />

feedback on ongoing research projects.<br />

Stakeholders and technical experts are being engaged through the following activities:<br />

• Technical roundtables with invited experts from diverse stakeholder groups to discuss the<br />

work underway to answer key research questions and identify possible topics for<br />

technical workshops. The roundtables also give the EPA access to a broad and balanced<br />

range of expertise as well as data from outside the agency.<br />

• Technical workshops with experts invited to participate in more in-depth discussions and<br />

share expertise on discrete technical topics relevant to the study.<br />

• Information requests through a Federal Register notice, requesting that the public submit<br />

relevant studies and data—particularly peer-reviewed studies—for the EPA’s<br />

consideration, including information on advances in industry practices and technologies.<br />

• Study updates to a wide range of stakeholders, including the general public, states, tribes,<br />

academia, non-governmental organizations, industry, professional organizations, and<br />

others.<br />

• Periodic briefings with the EPA’s Science Advisory Board to provide updates on the<br />

progress of the study.<br />

These efforts will help:<br />

• Inform the EPA’s interpretation of the research being conducted as part of this study.<br />

• Identify additional data and studies that may inform the report or results.<br />

• Identify future research needs.<br />

Additional information on the ongoing stakeholder engagement process can be found in Appendix B<br />

and online at http://www.epa.gov/<strong>hf</strong>study/. The website includes the presentations made by the<br />

EPA during the technical roundtables held in November 2012 as well as a list of roundtable<br />

participants. Readers are encouraged to check this website for up-to-date information on upcoming<br />

webinars for the general public and proceedings from technical workshops, which are currently<br />

scheduled for spring 2013.<br />

7


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

2. Overview of the Research Program<br />

The EPA’s Study of the Potential Impacts of Hydraulic Fracturing on Drinking Water Resources is<br />

organized into five topics according to the potential for interaction between hydraulic fracturing<br />

and drinking water resources. These five topics—stages of the hydraulic fracturing water cycle—<br />

are illustrated in Figure 1 and include (1) water acquisition, (2) chemical mixing, (3) well injection,<br />

(4) flowback and produced water, and (5) wastewater treatment and waste disposal.<br />

Figure 1. Illustration of the five stages of the hydraulic fracturing water cycle. The cycle includes the acquisition of<br />

water needed for the hydraulic fracturing fluid, onsite mixing of chemicals with the water to create the hydraulic<br />

fracturing fluid, injection of the fluid under high pressures to fracture the oil- or gas-containing formation, recovery of<br />

flowback and produced water (hydraulic fracturing wastewater) after the injection is complete, and treatment and/or<br />

disposal of the wastewater.<br />

Figure 2 lists potential drinking water issues identified for each stage of the hydraulic fracturing<br />

water cycle.<br />

8


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Water Use in Hydraulic<br />

Fracturing Operations<br />

Potential Drinking Water Issues<br />

Water Acquisition<br />

• Water availability<br />

• Impact of water withdrawal on water quality<br />

Chemical Mixing<br />

• Release to surface and ground water<br />

(e.g., onsite spills and/or leaks)<br />

• Chemical transportation accidents<br />

Well Injection<br />

• Accidental release to ground or surface water (e.g., well malfunction)<br />

• Fracturing fluid migration into drinking water aquifers<br />

• Formation fluid displacement into aquifers<br />

• Mobilization of subsurface formation materials into aquifers<br />

Flowback and<br />

Produced Water<br />

• Release to surface and ground water<br />

• Leakage from onsite storage into drinking water resources<br />

• Improper pit construction, maintenance, and/or closure<br />

Wastewater Treatment<br />

and Waste Disposal<br />

• Surface and/or subsurface discharge into surface and ground water<br />

• Incomplete treatment of wastewater and solid residuals<br />

• Wastewater transportation accidents<br />

Figure 2. Potential drinking water issues associated with each stage of the hydraulic fracturing water cycle. The potential issues help to define the fundamental<br />

research questions. Figure reprinted from the Study Plan (US EPA, 2011e).<br />

9


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

As described in the Study Plan, the potential issues led to the development of primary research<br />

questions that are supported by secondary research questions. The secondary research questions<br />

are addressed by the research projects listed in Table 1. Table 1 also provides short titles and<br />

descriptions of the research projects; these titles are used throughout the rest of the report.<br />

Table 1. Titles and descriptions of the research projects conducted as part of the EPA’s Study of the Potential<br />

Impacts of Hydraulic Fracturing on Drinking Water Resources. These titles are used throughout the rest of the report.<br />

Detailed descriptions of each project can be found in Chapters 3 through 7.<br />

Research Project<br />

Literature Review<br />

Spills Database Analysis<br />

Service Company Analysis<br />

Well File Review<br />

FracFocus Analysis<br />

Subsurface Migration Modeling<br />

Surface Water Modeling<br />

Water Availability Modeling<br />

Source Apportionment Studies<br />

Wastewater Treatability<br />

Studies<br />

Br-DBP Precursor Studies<br />

Analytical Method<br />

Development<br />

Description<br />

Analysis of Existing Data<br />

Review and assessment of existing papers and reports, focusing on<br />

peer-reviewed literature<br />

Analysis of selected federal and state databases for information on<br />

spills of hydraulic fracturing fluids and wastewaters<br />

Analysis of information provided by nine hydraulic fracturing service<br />

companies in response to a September 2010 information request on<br />

hydraulic fracturing operations<br />

Analysis of information provided by nine oil and gas operators in<br />

response to an August 2011 information request for 350 well files<br />

Analysis of data compiled from FracFocus, the national hydraulic<br />

fracturing chemical registry operated by the Ground Water Protection<br />

Council and the Interstate Oil and Gas Compact Commission<br />

Scenario Evaluations<br />

Numerical modeling of subsurface fluid migration scenarios that<br />

explore the potential for gases and fluids to move from the fractured<br />

zone to drinking water aquifers<br />

Modeling of concentrations of selected chemicals at public water<br />

supplies downstream from wastewater treatment facilities that<br />

discharge treated hydraulic fracturing wastewater to surface waters<br />

Assessment and modeling of current and future scenarios exploring<br />

the impact of water usage for hydraulic fracturing on drinking water<br />

availability in the Upper Colorado River Basin and the Susquehanna<br />

River Basin<br />

Laboratory Studies<br />

Identification and quantification of the source(s) of high bromide and<br />

chloride concentrations at public water supply intakes downstream<br />

from wastewater treatment plants discharging treated hydraulic<br />

fracturing wastewater to surface waters<br />

Assessment of the efficacy of common wastewater treatment<br />

processes on removing selected chemicals found in hydraulic<br />

fracturing wastewater<br />

Assessment of the ability of bromide and brominated compounds<br />

present in hydraulic fracturing wastewater to form brominated<br />

disinfection byproducts (Br-DBPs) during drinking water treatment<br />

processes<br />

Development of analytical methods for selected chemicals found in<br />

hydraulic fracturing fluids or wastewater<br />

Table continued on next page<br />

10


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

Research Project<br />

Toxicity Assessment<br />

Retrospective Studies<br />

Las Animas and Huerfano<br />

Counties, Colorado<br />

Dunn County, North<br />

Dakota<br />

Bradford County,<br />

Pennsylvania<br />

Washington County,<br />

Pennsylvania<br />

Wise County, Texas<br />

Prospective Studies<br />

Description<br />

Toxicity Assessment<br />

Toxicity assessment of chemicals reportedly used in hydraulic<br />

fracturing fluids or found in hydraulic fracturing wastewater<br />

Case Studies<br />

Investigations of whether reported drinking water impacts may be<br />

associated with or caused by hydraulic fracturing activities<br />

Investigation of potential drinking water impacts from coalbed<br />

methane extraction in the Raton Basin<br />

Investigation of potential drinking water impacts from a well blowout<br />

during hydraulic fracturing for oil in the Bakken Shale<br />

Investigation of potential drinking water impacts from shale gas<br />

development in the Marcellus Shale<br />

Investigation of potential drinking water impacts from shale gas<br />

development in the Marcellus Shale<br />

Investigation of potential drinking water impacts from shale gas<br />

development in the Barnett Shale<br />

Investigation of potential impacts of hydraulic fracturing through<br />

collection of samples from a site before, during, and after well pad<br />

construction and hydraulic fracturing<br />

Each project has been designed to inform answers to one or more of the secondary research<br />

questions with multiple projects informing answers to each secondary research question. The<br />

answers to the secondary research questions will then inform answers to the primary research<br />

questions. Figure 3 illustrates the relationship between water cycle stage, primary and secondary<br />

research questions, and research projects.<br />

11


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Figure 3. Illustration of the structure of the EPA’s Study of the Potential Impacts of Hydraulic Fracturing on Drinking<br />

Water Resources. Results from multiple research projects may be used to inform answers to one secondary research<br />

question. Additionally, one research project may provide information to help answer multiple secondary research<br />

questions. Each research project falls under one type of research activity.<br />

2.1. Research Questions<br />

This section describes the activities that occur during each stage of the water cycle, potential<br />

drinking water issues, and primary research questions, which are listed in Figure 4. 7 It also<br />

introduces the secondary research questions and lists the associated research projects. This section<br />

is intended to offer a broad overview of the EPA’s study and direct the reader to further<br />

information in subsequent chapters of this progress report. Later chapters (Chapters 3 through 7)<br />

contain detailed information about the progress of individual research projects listed in Tables 2<br />

through 6 below.<br />

7 Additional information on the hydraulic fracturing water cycle stages and research questions can be found in the Study<br />

Plan.<br />

12


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Water Use in Hydraulic<br />

Fracturing Operations<br />

Fundamental Research Question<br />

Water Acquisition<br />

What are the possible impacts of large volume water withdrawals<br />

from ground and surface waters on drinking water resources<br />

Chemical Mixing<br />

What are the possible impacts of surface spills on or near well pads<br />

of hydraulic fracturing fluids on drinking water resources<br />

Well Injection<br />

What are the possible impacts of the injection and fracturing<br />

process on drinking water resources<br />

Flowback and<br />

Produced Water<br />

What are the possible impacts of surface spills on or near well pads<br />

of flowback and produced water on drinking water resources<br />

Wastewater Treatment<br />

and Waste Disposal<br />

What are the possible impacts of inadequate treatment of<br />

hydraulic fracturing wastewaters on drinking water resources<br />

Figure 4. Fundamental research questions posed for each stage of the hydraulic fracturing water cycle. Figure reprinted from the Study Plan (US EPA, 2011e).<br />

13


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

2.1.1. Water Acquisition: What are the possible impacts of large volume water<br />

withdrawals from ground and surface waters on drinking water resources<br />

Hydraulic fracturing fluids are usually water-based, with approximately 90% of the injected fluid<br />

composed of water (GWPC and ALL Consulting, 2009). Estimates of water needs per well have been<br />

reported to range from 65,000 gallons for coalbed methane (CBM) production up to 13 million<br />

gallons for shale gas production, depending on the characteristics of the formation being fractured<br />

and the design of the production well and fracturing operation (GWPC and ALL Consulting, 2009;<br />

Nicot et al., 2011). Five million gallons of water are equivalent to the water used by approximately<br />

50,000 people for one day. 8 The source of the water may vary, but is typically ground water, surface<br />

water, or treated wastewater, as illustrated in Figure 5. Industry trends suggest a recent shift to<br />

using treated and recycled produced water (or other treated wastewaters) as base fluids in<br />

hydraulic fracturing operations.<br />

Figure 5. Water acquisition. Water for hydraulic fracturing can be drawn from a variety of sources including surface<br />

water, ground water, treated wastewater generated during previous hydraulic fracturing operations, and other types of<br />

wastewater.<br />

The EPA is working to better characterize the amounts and sources of water currently being used<br />

for hydraulic fracturing operations, including recycled water, and how these withdrawals may<br />

impact local drinking water quality and availability. To that end, secondary research questions have<br />

been developed, as well as the research projects listed in Table 2.<br />

8 This assumes that the average American uses approximately 100 gallons of water per day. See http://www.epa.gov/<br />

watersense/pubs/indoor.html.<br />

14


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table 2. Secondary research questions and applicable research projects identified for the water acquisition stage of<br />

the hydraulic fracturing water cycle. The table also identifies the sections of this report that contain detailed<br />

information about the listed research projects.<br />

Secondary Research Questions Applicable Research Projects Section<br />

Literature Review 3.1<br />

How much water is used in hydraulic fracturing<br />

operations, and what are the sources of this water<br />

How might water withdrawals affect short- and longterm<br />

water availability in an area with hydraulic<br />

fracturing activity<br />

What are the possible impacts of water withdrawals<br />

for hydraulic fracturing operations on local water<br />

quality<br />

Service Company Analysis 3.3<br />

Well File Review 3.4<br />

FracFocus Analysis 3.5<br />

Water Availability Modeling 4.3<br />

Literature Review 3.1<br />

Water Availability Modeling 4.3<br />

Literature Review 3.1<br />

2.1.2. Chemical Mixing: What are the possible impacts of surface spills on or near well<br />

pads of hydraulic fracturing fluids on drinking water resources<br />

Once onsite, water is mixed with chemicals to create the hydraulic fracturing fluid that is pumped<br />

down the well, as illustrated in Figure 6. The fluid serves two purposes: to create pressure to<br />

propagate fractures and to carry the proppant into the fracture. Chemicals are added to the fluid to<br />

change its properties (e.g., viscosity, pH) in order to optimize the performance of the fluid. Roughly<br />

1% of water-based hydraulic fracturing fluids are composed of various chemicals, which is<br />

equivalent to 50,000 gallons for a shale gas well that uses 5 million gallons of fluid.<br />

Figure 6. Chemical mixing. Water is mixed with chemicals and proppant onsite to create the hydraulic fracturing fluid<br />

immediately before injection.<br />

15


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Hydraulic fracturing operations require large quantities of supplies, equipment, water, and vehicles.<br />

Onsite storage, mixing, and pumping of hydraulic fracturing fluids may result in accidental releases,<br />

such as spills or leaks. 9 Released fluids could then flow into nearby surface water bodies or<br />

infiltrate into the soil and near-surface ground water, potentially reaching drinking water<br />

resources. In order to explore the potential impacts of surface releases of hydraulic fracturing fluids<br />

on drinking water resources, the EPA is: (1) compiling information on reported spills; (2)<br />

identifying chemical additives used in hydraulic fracturing fluids and their chemical, physical, and<br />

toxicological properties; and (3) gathering data on the environmental fate and transport of selected<br />

hydraulic fracturing chemical additives. These activities correspond to the secondary research<br />

questions and research projects described in Table 3.<br />

Table 3. Secondary research questions and applicable research projects identified for the chemical mixing stage of<br />

the hydraulic fracturing water cycle. The table also identifies the sections of this report that contain detailed<br />

information about the listed research projects.<br />

Secondary Research Questions Applicable Research Projects Section<br />

What is currently known about the frequency, severity,<br />

and causes of spills of hydraulic fracturing fluids and<br />

additives<br />

What are the identities and volumes of chemicals<br />

used in hydraulic fracturing fluids, and how might this<br />

composition vary at a given site and across the<br />

country<br />

What are the chemical, physical, and toxicological<br />

properties of hydraulic fracturing chemical additives<br />

If spills occur, how might hydraulic fracturing chemical<br />

additives contaminate drinking water resources<br />

Literature Review 3.1<br />

Spills Database Analysis 3.2<br />

Service Company Analysis 3.3<br />

Well File Review 3.4<br />

Literature Review 3.1<br />

Service Company Analysis 3.3<br />

FracFocus Analysis 3.5<br />

Analytical Method Development 5.4<br />

Toxicity Assessment 6<br />

Literature Review 3.1<br />

Retrospective Case Studies 7<br />

2.1.3. Well Injection: What are the possible impacts of the injection and fracturing<br />

process on drinking water resources<br />

The hydraulic fracturing fluid is pumped down the well at pressures great enough to fracture the<br />

oil- or gas-containing rock formation, as shown in Figure 7 for both horizontal and vertical well<br />

completions. Production wells are drilled and completed in order to best and most efficiently drain<br />

the geological reservoir of its hydrocarbon resources. This means that wells may be drilled and<br />

completed vertically (panel b in Figure 7), vertically at the top and then horizontally at the bottom<br />

(panel a), or in other configurations deviating from vertical, known as “deviated wells.”<br />

9 As noted in the Study Plan, transportation-related spills of hydraulic fracturing chemical additives and wastewater are<br />

outside of the scope of the current study.<br />

16


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

(a)<br />

(b)<br />

Figure 7. Well injection. During injection, hydraulic fracturing fluids are pumped into the well at high pressures, which<br />

are sustained until the fractures are formed. Hydraulic fracturing can be used with both (a) deep, horizontal well<br />

completions and (b) shallower, vertical well completions. Horizontal wells are typically used in formations such as<br />

tight sandstones, carbonate rock, and shales. Vertical wells are typically used in formations for conventional<br />

production and coalbed methane.<br />

Within this stage of the hydraulic fracturing water cycle, the EPA is studying a number of scenarios<br />

that may lead to changes in local drinking water resources, including well construction failure and<br />

induced fractures intersecting existing natural (e.g., faults or fractures) or man-made (e.g.,<br />

abandoned wells) features that may act as conduits for contaminant transport. Table 4 lists the<br />

secondary research questions and research projects that address these concerns.<br />

Table 4. Secondary research questions and applicable research projects identified for the well injection stage of the<br />

hydraulic fracturing water cycle. The table also identifies the sections of this report that contain detailed information<br />

about the listed research projects.<br />

Secondary Research Questions Applicable Research Projects Section<br />

Literature Review 3.1<br />

How effective are current well construction practices Service Company Analysis 3.3<br />

at containing gases and fluids before, during, and Well File Review 3.4<br />

after fracturing<br />

Subsurface Migration Modeling 4.1<br />

Retrospective Case Studies 7<br />

Literature Review 3.1<br />

Can subsurface migration of fluids or gases to<br />

drinking water resources occur, and what local<br />

geologic or man-made features might allow this<br />

Service Company Analysis 3.3<br />

Well File Review 3.4<br />

Subsurface Migration Modeling 4.1<br />

Retrospective Case Studies 7<br />

17


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

2.1.4. Flowback and Produced Water: What are the possible impacts of surface spills on<br />

or near well pads of flowback and produced water on drinking water resources<br />

When the injection pressure is reduced, the direction of fluid flow reverses, leading to the recovery<br />

of flowback and produced water. For this study, “flowback” is the fluid returned to the surface after<br />

hydraulic fracturing has occurred, but before the well is placed into production, while “produced<br />

water” is the fluid returned to the surface after the well has been placed into production. 10 They are<br />

collectively referred to as “hydraulic fracturing wastewater” and may contain chemicals injected as<br />

part of the hydraulic fracturing fluid, substances naturally occurring in the oil- or gas-producing<br />

formation, 11 hydrocarbons, and potential reaction and degradation products.<br />

Figure 8. Flowback and produced water. During this stage, the pressure on the hydraulic fracturing fluid is reduced<br />

and the flow is reversed. The flowback and produced water contain hydraulic fracturing fluids, native formation water,<br />

and a variety of naturally occurring substances picked up by the wastewater during the fracturing process. The fluids<br />

are separated from any gas or oil produced with the water and stored in either tanks or an open pit.<br />

As depicted in Figure 8, the wastewater is typically stored onsite in impoundment pits or tanks.<br />

Onsite transfer and storage of hydraulic fracturing wastewater may result in accidental releases,<br />

such as spills or leaks, which may reach nearby drinking water resources. The potential impacts to<br />

drinking water resources from flowback and produced water are similar to the potential impacts<br />

identified in the chemical mixing stage of the hydraulic fracturing water cycle, with the exception of<br />

different fluid compositions for injected fluids and wastewater. Therefore, the secondary research<br />

10 Produced water is a product of all oil and gas wells, including wells that have not been hydraulically fractured.<br />

11 Substances naturally found in hydraulically fractured formations may include brines, trace elements (e.g., mercury,<br />

lead, arsenic), naturally occurring radioactive material (e.g., radium, thorium, uranium), gases (e.g., natural gas, hydrogen<br />

sulfide), and organic material (e.g., organic acids, polycyclic aromatic hydrocarbons, volatile organic compounds).<br />

18


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

questions and associated research projects are similar. The secondary research questions and<br />

applicable research projects are listed in Table 5.<br />

Table 5. Secondary research questions and applicable research projects identified for the flowback and produced<br />

water stage of the hydraulic fracturing water cycle. The table also identifies the sections of this report that contain<br />

detailed information about the listed research projects.<br />

Secondary Research Questions Applicable Research Projects Section<br />

Literature Review 3.1<br />

What is currently known about the frequency, severity, Spills Database Analysis 3.2<br />

and causes of spills of flowback and produced water Service Company Analysis 3.3<br />

Well File Review 3.4<br />

What is the composition of hydraulic fracturing<br />

wastewaters, and what factors might influence this<br />

composition<br />

What are the chemical, physical, and toxicological<br />

properties of hydraulic fracturing wastewater<br />

constituents<br />

If spills occur, how might hydraulic fracturing<br />

wastewater contaminate drinking water resources<br />

Literature Review 3.1<br />

Service Company Analysis 3.3<br />

Well File Review 3.4<br />

Analytical Method Development 5.4<br />

Toxicity Assessment 6<br />

Literature Review 3.1<br />

Retrospective Case Studies 7<br />

2.1.5. Wastewater Treatment and Waste Disposal: What are the possible impacts of<br />

inadequate treatment of hydraulic fracturing wastewaters on drinking water<br />

resources<br />

Estimates of the fraction of hydraulic fracturing wastewater recovered vary by geologic formation<br />

and range from 10% to 70% of the injected hydraulic fracturing fluid (GWPC and ALL Consulting,<br />

2009; US EPA, 2011f). For a hydraulic fracturing job that uses 5 million gallons of hydraulic<br />

fracturing fluid, this means that between 500,000 and 3.5 million gallons of fluid will be returned to<br />

the surface. As illustrated in Figure 9, the wastewater is generally managed through disposal into<br />

deep underground injection control (UIC) wells, 12 treatment followed by discharge to surface water<br />

bodies, 13 or treatment followed by reuse.<br />

12 Underground injection of fluids related to oil and gas production (including flowback and produced water) is<br />

authorized by the Safe Drinking Water Act.<br />

13 Treatment processes involving discharge to surface waters are authorized by the Clean Water Act and the National<br />

Pollutant Discharge Elimination System program.<br />

19


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Figure 9. Wastewater treatment and waste disposal. Flowback and produced water is frequently disposed of in deep<br />

injection wells, but may also be trucked, or in some cases piped, to a disposal or recycling facility. Once treated, the<br />

wastewater may be reused in subsequent hydraulic fracturing operations or discharged to surface water.<br />

Understanding the treatment, disposal, and reuse of flowback and produced water from hydraulic<br />

fracturing activities is important. For example, contaminants present in these waters may be<br />

inadequately treated at publicly owned treatment works (POTWs), discharges from which may<br />

threaten downstream drinking water intakes, as depicted in Figure 9. 14 Table 6 summarizes the<br />

secondary research questions and the applicable research projects for each question.<br />

14 As noted in the Study Plan, this study does not propose to evaluate the potential impacts of underground injection or<br />

the associated potential impacts due to transport and storage leading up to ultimate disposal in a UIC well.<br />

20


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table 6. Secondary research questions and applicable research projects identified for the wastewater treatment and<br />

waste disposal stage of the hydraulic fracturing water cycle. The table also identifies the sections of this report that<br />

contain detailed information about the listed research projects.<br />

Secondary Research Questions Applicable Research Projects Section<br />

What are the common treatment and disposal Literature Review 3.1<br />

methods for hydraulic fracturing wastewater, and Well File Review 3.4<br />

where are these methods practiced<br />

FracFocus Analysis 3.5<br />

How effective are conventional POTWs and<br />

Literature Review 3.1<br />

commercial treatment systems in removing organic<br />

and inorganic contaminants of concern in hydraulic Wastewater Treatability Studies 5.2<br />

fracturing wastewater<br />

Literature Review 3.1<br />

What are the potential impacts from surface water<br />

Surface Water Modeling 4.2<br />

disposal of treated hydraulic fracturing wastewater on<br />

drinking water treatment facilities<br />

Source Apportionment Studies 5.1<br />

Br-DBP Precursor Studies 5.3<br />

2.2. Environmental Justice<br />

Environmental justice is the fair treatment and meaningful involvement of all people regardless of<br />

race, color, national origin, or income, with respect to the development, implementation, and<br />

enforcement of environmental laws, regulations, and policies. 15<br />

During the planning process, some stakeholders raised concerns about environmental justice and<br />

hydraulic fracturing, while others stated that hydraulic fracturing–related activities provide<br />

benefits to local communities. In its review of the draft Study Plan, the EPA’s Science Advisory<br />

Board supported the inclusion in the study of an environmental justice analysis as it pertains to the<br />

potential impacts on drinking water resources. The EPA, therefore, attempted to conduct a<br />

screening to provide insight into the research questions in Table 7.<br />

Table 7. Research questions addressed by assessing the demographics of locations where hydraulic fracturing<br />

activities are underway.<br />

Fundamental Research Question<br />

Does hydraulic fracturing<br />

disproportionately occur in or near<br />

communities with environmental<br />

justice concerns<br />

Secondary Research Questions<br />

• Are large volumes of water being disproportionately<br />

withdrawn from drinking water resources that serve<br />

communities with environmental justice concerns<br />

• Are hydraulically fractured oil and gas wells<br />

disproportionately located near communities with<br />

environmental justice concerns<br />

• Is wastewater from hydraulic fracturing operations being<br />

disproportionately treated or disposed of (via POTWs or<br />

commercial treatment systems) in or near communities with<br />

environmental justice concerns<br />

15 The EPA’s definition of environmental justice can be found at<br />

http://www.epa.gov/environmentaljustice/basics/index.html and was informed by E.O. 12898.<br />

21


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Environmental justice screening uses easily obtained environmental and demographic information<br />

to highlight locations where additional review (i.e., information collection or analysis) may be<br />

warranted (US EPA, 2012c). Screenings do not examine whether co-location of specific activities<br />

and communities with certain demographics (e.g., low-income, non-white minority, young children,<br />

and elderly subpopulations) may lead to any positive or negative impacts on a given community.<br />

Nationwide data on the locations of water withdrawals and wastewater treatment associated with<br />

hydraulic fracturing activities are difficult to obtain. The EPA was not able to identify<br />

comprehensive data sources that identify the locations of water withdrawals associated with<br />

hydraulic fracturing or facilities receiving hydraulic fracturing wastewaters. Geographic data on<br />

hydraulic fracturing-only water use (rather than general oil and gas water use) are limited, and the<br />

available data are aggregated by regions too large for an environmental justice analysis. Data on<br />

commercial and publicly owned treatment works accepting hydraulic fracturing wastewater were<br />

found to be inconsistent between states or difficult to obtain.<br />

Data on the locations of hydraulically fractured oil and gas production wells considered for the<br />

environmental justice screen are available from two sources: data provided to the EPA from nine<br />

hydraulic fracturing service companies (see Section 3.3) and data obtained from FracFocus (Section<br />

3.5). The service company data set includes county-level locations of approximately 25,000 oil and<br />

gas wells hydraulically fractured between September 2009 and October 2010. In total, 590 of the<br />

3,221 counties in the United States contained wells hydraulically fractured by the nine service<br />

companies during the period under analysis. In comparison, the FracFocus data set includes<br />

latitude/longitude and county-level information on the location of roughly 11,000 wells<br />

hydraulically fractured between January 2009 and February 2012. In total, only 251 of the 3,221<br />

counties in the United States contained wells reported to FracFocus during this time period.<br />

The county-level resolution provided by the service company data set is insufficient for<br />

determining whether hydraulic fracturing activities are occurring in communities that possess<br />

characteristics associated with environmental justice populations. Finer resolution is needed since<br />

counties can contain a multitude of communities, townships, and even cities, with diverse<br />

populations. Data obtained from FracFocus provide well locations at finer resolution (i.e., specific<br />

latitude/longitude coordinates), which may provide further opportunity for either state- or<br />

nationwide environmental justice screens.<br />

2.3. Changes to the Research Program<br />

The EPA has significantly modified some of the research projects since the publication of the Study<br />

Plan. These modifications are discussed below.<br />

FracFocus Analysis. In early 2011, the Ground Water Protection Council and the Interstate Oil and<br />

Gas Compact Commission jointly launched a new national registry for chemicals used in hydraulic<br />

fracturing, called FracFocus. This registry is an online repository where oil and gas well operators<br />

can upload information regarding the chemical composition of hydraulic fracturing fluids used in<br />

specific oil and gas production wells. Extracting data from FracFocus allows the EPA to gather<br />

publicly available, nationwide data on the water volumes and chemicals used in hydraulic<br />

fracturing operations, as reported by oil and gas operating companies. These data are being<br />

22


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

analyzed to identify chemicals used in hydraulic fracturing fluids as well as the geographic<br />

distribution of water and chemical use.<br />

Prospective Case Studies. The EPA identified the location of one of the prospective case studies as De<br />

Soto Parish, Louisiana, in the Haynesville Shale. Due to scheduling conflicts, the location in De Soto<br />

Parish is no longer being considered for a prospective case study.<br />

The EPA continues to work with industry partners to identify locations and develop research<br />

activities for prospective case studies. As part of these case studies, the EPA intends to monitor<br />

local water quality for up to a year or more after hydraulic fracturing occurs. It is likely, therefore,<br />

that the prospective case studies will be completed after the report of results. In that event, results<br />

from any prospective case studies will be published in a follow-up report.<br />

Chemical Prioritization. As part of the toxicity assessment research project, the EPA is compiling<br />

chemical, physical, and toxicological properties for chemicals reportedly used in hydraulic<br />

fracturing fluids and/or detected in flowback and produced water. One aspect of the planned<br />

second phase of this work was to include prioritizing a subset of these chemicals for future toxicity<br />

screening using high throughput screening assays. However, consistent with recommendations of<br />

the Science Advisory Board, the agency will not conduct high throughput screening assays at this<br />

time on a subset of these chemicals, but will continue efforts to identify, evaluate, and prioritize<br />

existing toxicity data.<br />

Reactions Between Hydraulic Fracturing Fluids and Shale. Based on research already being<br />

conducted by the US Department of Energy and academic institutions on the interactions between<br />

hydraulic fracturing fluids and various rock formations, 16 the EPA has decided to discontinue its<br />

work in this area. The EPA continues to believe in the importance of research to address research<br />

questions associated with this project, but has decided to rely upon work being conducted by<br />

another federal agency.<br />

Therefore, the EPA has removed two research questions associated with this project:<br />

• How might hydraulic fracturing fluids change the fate and transport of substances in the<br />

subsurface through geochemical interactions<br />

• What are the chemical, physical, and toxicological properties of substances in the<br />

subsurface that may be released by hydraulic fracturing operations<br />

2.4. Research Approach<br />

The research projects listed in Table 1 and discussed in detail in Chapters 3 through 7 of this<br />

progress report require a broad range of scientific expertise in environmental and petroleum<br />

engineering, ground water hydrology, fate and transport modeling, and toxicology, as well as many<br />

other disciplines. Consequently, the EPA is using a transdisciplinary research approach that<br />

16 See, for example, research underway by the US Department of Energy’s National Energy Technology Laboratory<br />

(http://www.netl.doe.gov/publications/factsheets/rd/R%26D166.pdf) and Penn State 3S Laboratory<br />

(http://3s.ems.psu.edu/research.html).<br />

23


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

integrates various types of expertise from inside and outside the agency. The research projects fall<br />

into five categories: analysis of existing data, case studies, scenario modeling and evaluation,<br />

laboratory studies, and toxicology assessments. Table 8 summarizes the five main types of research<br />

activities occurring as part of this study and their objectives. Figure 3 illustrates the relationship<br />

between the research activities and the research projects and questions.<br />

Table 8. Research activities and objectives. Each research project falls under one type of research activity.<br />

Activity<br />

Analysis of existing data<br />

Scenario evaluations<br />

Laboratory studies<br />

Toxicity assessment<br />

Case studies<br />

Retrospective<br />

Prospective<br />

Objective<br />

Gather and summarize existing data from various sources to provide<br />

current information on hydraulic fracturing activities; includes information<br />

requested of hydraulic fracturing service companies and oil and gas<br />

operators*<br />

Use computer modeling to assess the potential for hydraulic fracturing to<br />

impact drinking water resources<br />

Conduct targeted experiments to test and develop analytical detection<br />

methods and to study the fate and transport of selected chemicals during<br />

wastewater treatment and discharge to surface water<br />

Identify chemicals used in hydraulic fracturing fluids or reported to be in<br />

hydraulic fracturing wastewater and compile available chemical, physical,<br />

and toxicological properties<br />

Study sites with reported contamination to understand the underlying<br />

causes and potential impacts to drinking water resources<br />

Develop understanding of hydraulic fracturing processes and their<br />

potential impacts on drinking water resources<br />

* For more information on the information requests, see http://www.epa.gov/<strong>hf</strong>study/analysis-of-existing-data.html.<br />

24


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

3. Analysis of Existing Data<br />

The objective of this approach is to gather and summarize data from many sources to provide<br />

current information on hydraulic fracturing activities. The EPA is collecting and analyzing data on<br />

chemical spills, surface water discharges, and chemicals found in hydraulic fracturing fluids and<br />

wastewater, among others. These data have been collected from a variety of sources, including state<br />

and federal agencies, industry, and public sources. Included among these sources is information<br />

received after the September 2010 letter requesting data from nine hydraulic fracturing service<br />

companies and the August 2011 letter requesting well files from nine oil and gas well operators. 17<br />

This chapter includes progress reports for the following projects:<br />

3.1. Literature Review.................................................................................................................................................. 25<br />

Review and assessment of existing papers and reports, focusing on peer-reviewed literature<br />

3.2. Spills Database Analysis...................................................................................................................................... 31<br />

Analysis of selected federal and state databases for information on spills of hydraulic<br />

fracturing fluids and wastewaters<br />

3.3. Service Company Analysis ................................................................................................................................. 39<br />

Analysis of information provided by nine hydraulic fracturing service companies in response to<br />

a September 2010 information request on hydraulic fracturing operations<br />

3.4. Well File Review..................................................................................................................................................... 46<br />

Analysis of information provided by nine oil and gas operators in response to an August 2011<br />

information request for 350 well files<br />

3.5. FracFocus Analysis................................................................................................................................................ 54<br />

Analysis of data compiled from FracFocus, the national hydraulic fracturing chemical registry<br />

operated by the Ground Water Protection Council and the Interstate Oil and Gas Compact<br />

Commission<br />

3.1. Literature Review<br />

3.1.1. Relationship to the Study<br />

The EPA is gathering and assessing literature relevant to all secondary research questions.<br />

3.1.2. Project Introduction<br />

An extensive review of existing literature is an important component of the EPA’s study of the<br />

relationship between hydraulic fracturing and drinking water resources. The objective of this<br />

literature review is to identify and analyze data and literature relevant to all secondary research<br />

questions. This objective will be met by reviewing a wide range of information sources on the five<br />

stages of the hydraulic fracturing water cycle. Sources identified through the literature review are<br />

subject to a quality review to support decisions regarding their inclusion in the EPA’s report of<br />

17 Copies of these information requests are available at http://www.epa.gov/<strong>hf</strong>study/analysis-of-existing-data.html.<br />

25


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

results. Information gathered during the literature review will be synthesized with results from the<br />

other research projects described in this progress report to answer the research questions posed in<br />

the Study Plan and summarized in Chapter 2.<br />

3.1.3. Research Approach<br />

Existing literature and data is being identified through a variety of methods, including conducting a<br />

search of published documents, searching online databases such as OnePetro 18 and Web of<br />

Knowledge SM 19 and reviewing materials provided to the EPA through technical workshops,<br />

comment submissions, and the Science Advisory Board’s review of the draft study plan. 20 Once<br />

identified, sources are classified as shown in Table 9.<br />

Table 9. Classifications of information sources with examples. Once identified, existing literature and data sources<br />

are classified according to the following categories.<br />

Source Classification Examples<br />

Journal publications, reports, and white papers developed by federal and<br />

Peer-reviewed literature<br />

state agencies<br />

Non-peer-reviewed<br />

literature<br />

Unpublished data<br />

Non-peer-reviewed government documents; congressional documents<br />

and hearing proceedings; workshop proceedings; Ph.D. theses; nonpeer-reviewed<br />

reports and white papers from industry, associations, and<br />

non-governmental organizations<br />

Online databases, personal communications, unpublished manuscripts,<br />

unpublished government data<br />

Once sources are grouped into the categories shown in Table 9 above, assessment factors are used<br />

to further evaluate their merit. Five assessment factors are being used to evaluate the quality of<br />

existing data and information: soundness, applicability and utility, clarity and completeness,<br />

uncertainty and variability, and evaluation and review (US EPA, 2003a). These factors are described<br />

in more detail in Table 10.<br />

18 OnePetro is an online library of technical literature for the oil and gas exploration and production industry. It can be<br />

accessed at http://www.onepetro.org/.<br />

19 Thomson Reuters Web of Knowledge SM is a research platform that provides access to objective content and powerful<br />

tools to search, track, measure, and collaborate in the sciences, social sciences, arts, and humanities. It can be accessed at<br />

http://wokinfo.com/.<br />

20 A list of literature recommended by the Science Advisory Board can be found on pages 29–34 of the Science Advisory<br />

Board’s review of the draft Study Plan, available at http://yosemite.epa.gov/sab/sabproduct.nsf/0/<br />

2BC3CD632FCC0E99852578E2006DF890/$File/EPA-SAB-11-012-unsigned.pdf.<br />

26


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table 10. Description of factors used to assess the quality of existing data and information compiled during the<br />

literature review. The assessment factors are identified in (US EPA, 2003a).<br />

Factors<br />

Soundness<br />

Applicability and utility<br />

Clarity and<br />

completeness<br />

Uncertainty and<br />

variability<br />

Evaluation and review<br />

Description<br />

The extent to which the scientific and technical procedures, measures,<br />

methods, or models employed to generate the information are reasonable<br />

for, and consistent with, the intended application<br />

The extent to which the information is relevant for the agency’s intended use<br />

The degree of clarity and completeness with which the data, assumptions,<br />

methods, quality assurance, sponsoring organizations, and analyses<br />

employed to generate the information are documented<br />

The extent to which the variability and uncertainty (quantitative and<br />

qualitative) in the information or in the procedures, measures, methods or<br />

models are evaluated and characterized<br />

The extent of independent verification, validation, and peer review of the<br />

information or of the procedures, measures, methods, or models<br />

Information included in the report of results will be drawn primarily from peer-reviewed<br />

publications. Peer-reviewed publications contain the most reliable information, although some<br />

portions of the report may contain compilations of data from a variety of sources and source<br />

classifications. Non-peer-reviewed and unpublished sources will not form the sole basis of any<br />

conclusions presented in the report of results. Generally, these sources will be used to support<br />

results presented from peer-reviewed work, enhance understanding based on peer-reviewed<br />

sources, identify promising ideas of investigation, and discuss further in-depth work needed.<br />

The criteria in Table 10 are applied to all sources to ensure that the EPA is using high-quality data.<br />

In some cases, these data may not strictly meet the quality guidelines outlined in Table 10, though<br />

they still provide valuable information. Principal investigators on this project are responsible for<br />

deciding whether to include these data and providing all available background information in order<br />

to place these results in the appropriate context.<br />

3.1.4. Status and Preliminary Data<br />

The literature review is currently underway. Water acquisition, chemical mixing, and flowback and<br />

produced water are the only stages of the hydraulic fracturing water cycle for which specific<br />

updates are available at this time.<br />

Water Acquisition. The water acquisition literature review is intended to complement the analysis<br />

of existing data on hydraulic fracturing fluid source water resources from nine service companies<br />

(see Section 3.3) and nine oil and gas operators (Section 3.4), as well as the analysis of existing data<br />

from FracFocus (Section 3.5). Work at this stage is directed at answering three secondary research<br />

questions:<br />

• How much water is used in hydraulic fracturing operations, and what are the sources of<br />

this water<br />

• How might water withdrawals affect short- and long-term water availability in an area with<br />

hydraulic fracturing activity<br />

27


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

• What are the possible impacts of water withdrawals for hydraulic fracturing operations on<br />

local water quality<br />

To date, work has focused on the first question regarding the volumes and sources of water<br />

acquired for use in hydraulic fracturing. The literature review focuses on the major basins where<br />

hydraulic fracturing is prevalent in order to present a national perspective on water use.<br />

Hydrocarbon plays that will be highlighted include the Barnett, Eagle Ford, and Haynesville Shales<br />

in the South, the Bakken Shale in the Midwest, and the Marcellus and Utica Shales in the East.<br />

The Barnett, Eagle Ford, and Haynesville Shales have undergone the most thorough analysis as<br />

reflected by the availability of peer-reviewed literature pertaining to the Texas oil and gas basins<br />

and to the water resources in the southern United States. The Bakken Shale has also been<br />

investigated extensively, although very little peer-reviewed literature was available for analysis as<br />

of July 2012. Instead, information on volumes and sources of water in the Bakken Shale comes<br />

largely from news articles. Water acquisition in the Marcellus and Utica Shales has not yet been<br />

analyzed, but water withdrawal data is expected to be available.<br />

Chemical Mixing and Flowback and Produced Water. Existing scientific literature is being reviewed<br />

to identify how chemicals used in hydraulic fracturing fluids or present in hydraulic fracturing<br />

wastewaters may contaminate drinking water resources as a result of surface spills of these fluids.<br />

Relevant information from the literature review will help address the research questions listed<br />

below:<br />

• If spills occur, how might hydraulic fracturing chemical additives contaminate drinking<br />

water resources<br />

• If spills occur, how might hydraulic fracturing wastewaters contaminate drinking water<br />

resources<br />

The EPA has identified chemicals for further review based on publicly available information on<br />

hazard and frequency of use. Tables 11 and 12 identify a subset of chemicals used in hydraulic<br />

fracturing fluids as reported to the US House of Representatives’ Committee on Energy and<br />

Commerce by 14 hydraulic fracturing service companies as being used in hydraulic fracturing fluids<br />

between 2005 and 2009 (USHR, 2011). Table 11 lists chemicals that are known or suspected<br />

carcinogens, regulated by the Safe Drinking Water Act (SDWA), or listed as Clean Air Act hazardous<br />

air pollutants. The Committee included the hazardous air pollutant designation for listed chemicals<br />

because some may impact drinking water (e.g., methanol and ethylene glycol). Table 12 lists the<br />

chemical components appearing most often in over 2,500 hydraulic fracturing products used<br />

between 2005 and 2009, according to the information reported to the Committee.<br />

28


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table 11. Chemicals identified by the US House of Representatives Committee on Energy and Commerce as known<br />

or suspected carcinogens, regulated under the Safe Drinking Water Act (SDWA) or classified as hazardous air<br />

pollutants (HAP) under the Clean Air Act. The number of products containing each chemical is also listed. These<br />

chemicals were reported by 14 hydraulic fracturing service companies to be in a total of 652 different products used<br />

between 2005 and 2009. Reproduced from USHR (2011).<br />

Chemicals Category No. of Products<br />

Methanol HAP 342<br />

Ethylene glycol HAP 119<br />

Naphthalene Carcinogen, HAP 44<br />

Xylene SDWA, HAP 44<br />

Hydrochloric acid HAP 42<br />

Toluene SDWA, HAP 29<br />

Ethylbenzene SDWA, HAP 28<br />

Diethanolamine HAP 14<br />

Formaldehyde Carcinogen, HAP 12<br />

Thiourea Carcinogen 9<br />

Benzyl chloride Carcinogen, HAP 8<br />

Cumene HAP 6<br />

Nitrilotriacetic acid Carcinogen 6<br />

Dimethyl formamide HAP 5<br />

Phenol HAP 5<br />

Benzene Carcinogen, SDWA, HAP 3<br />

Di (2-ethylhexyl) phthalate Carcinogen, SDWA, HAP 3<br />

Acrylamide Carcinogen, SDWA, HAP 2<br />

Hydrofluoric acid HAP 2<br />

Phthalic anhydride HAP 2<br />

Acetaldehyde Carcinogen, HAP 1<br />

Acetophenone HAP 1<br />

Copper SDWA 1<br />

Ethylene oxide Carcinogen, HAP 1<br />

Lead Carcinogen, SDWA, HAP 1<br />

Propylene oxide Carcinogen, HAP 1<br />

p-Xylene HAP 1<br />

Table 12. Chemical appearing most often in hydraulic fracturing in over 2,500 products reported by 14 hydraulic<br />

fracturing service companies as being used between 2005 and 2009. Reproduced from USHR (2011).<br />

Chemical<br />

No. of Products<br />

Methanol 342<br />

Isopropanol 274<br />

Crystalline silica 207<br />

2-Butoxyethanol 126<br />

Ethylene glycol 119<br />

Hydrotreated light petroleum distillates 89<br />

Sodium hydroxide 80<br />

29


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Existing scientific literature is also being reviewed for the chemicals identified as part of the<br />

analytical method development research project (see Table 45 in Section 5.4). This table includes<br />

chemicals associated with injected hydraulic fracturing fluids and wastewater.<br />

Literature searches have found papers describing impacts from spills of produced water (Healy et<br />

al., 2011; Healy et al., 2008), although the emphasis is often on ecosystem impacts rather than<br />

drinking water impacts. Produced water has the greatest number of literature publications for<br />

reported spills compared to hydraulic fracturing fluids and flowback, because produced water must<br />

be managed in both conventional and unconventional oil and gas production. Papers describing<br />

impacts from spills of produced water from conventional oil and gas production wells are being<br />

considered as part of the literature review because the chemical composition of flowback and<br />

produced water from hydraulically fractured formations is similar to that of conventional<br />

reservoirs (Hayes, 2009). Publications about impoundment leaks or other types of surface<br />

impoundment failures are also included within the scope of the flowback and produced water<br />

literature review.<br />

Because some of the chemicals commonly used in hydraulic fracturing fluid are ubiquitous, a very<br />

large numbers of papers have been found. To narrow the scope, recent review papers on<br />

environmental impacts and other published summaries on transport of chemicals or classes of<br />

chemicals are being sought. Information on the chemicals listed in Tables 11, 12, and 45 has been<br />

collected primarily by searching peer-reviewed literature using keyword searches of major<br />

databases, including Web of Knowledge SM , Proquest, 21 and OnePetro. Review papers describing<br />

impacts from spills of hydraulic fracturing fluids containing benzene, toluene, ethylbenzene, and<br />

xylenes (Farhadian et al., 2008; Seagren and Becker, 2002; Seo et al., 2009); ethylene glycol<br />

(Staples et al., 2001); phenol (Van Schie and Young L.Y., 2000); surfactants (Scott and Jones, 2000;<br />

Sharma et al., 2009; Soares A. et al., 2008; Van Ginkel, 1996); and napthalenes (Haritash and<br />

Kaushik, 2009; Rogers et al., 2002) have been identified. Other sources of information include the<br />

Government Accountability Office report on federal research on produced water (US GAO, 2012);<br />

toxicological profiles from the Agency for Toxic Substances and Disease Registry, which often<br />

contain brief summaries of information on transport and transformation; 22 EPA software systems<br />

(US EPA, 2012b); and chemical reference handbooks (Howard, 1989; Howard et al., 1991;<br />

Montgomery, 2000). Specific discussion of abiotic transformations is included in some of these<br />

references, including the Agency for Toxic Substances and Disease Registry Toxicological Profiles,<br />

environmental organic chemistry references (Schwarzenbach et al., 2002), and review papers<br />

(Stangroom et al., 2010).<br />

Chemical and physical properties of most of the organic chemicals listed in Tables 11 and 12 have<br />

been summarized, and the analysis is nearly complete. As more chemicals of interest are identified<br />

throughout the study, the number of chemicals may expand. Fewer publications exist for less<br />

21 ProQuest can be accessed at http://www.proquest.com.<br />

22 See, for example, pages 258–259 of ATSDR (2007).<br />

30


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

common chemicals, however, and obtaining enough data to characterize these chemicals’ potential<br />

to affect drinking water resources may not be feasible.<br />

3.1.5. Next Steps<br />

Next steps include completing the literature review for questions pertaining to sources, volumes,<br />

and impacts of large volume water withdrawals on local water quality and water availability.<br />

Further review of the water acquisition and quantity literature will specifically address the volumes<br />

and sources of water used in the Marcellus and Utica Shales. The literature review on chemical<br />

mixing and flowback and produced water for information that may answer the secondary research<br />

questions for those water stages will be completed. The EPA will also review relevant literature on<br />

all the remaining secondary research questions.<br />

3.1.6. Quality Assurance Summary<br />

The quality assurance project plan (QAPP) for the literature review, “Data and Literature<br />

Evaluation for the EPA’s Study of the Potential Impacts of Hydraulic Fracturing (HF) on Drinking<br />

Water Resources (Version 0),” was approved on September 4, 2012 (US EPA, 2012f). Links to the all<br />

of the QAPPs are provided in Appendix C.<br />

3.2. Spills Database Analysis<br />

3.2.1. Relationship to the Study<br />

The primary research questions for the chemical mixing and flowback and produced water stages<br />

of the hydraulic fracturing water cycle focus on the potential for hydraulic fracturing fluids and<br />

wastewaters to be spilled on the surface, possibly impacting nearby drinking water resources. This<br />

project searches various data sources in order to answer the research questions listed in Table 13.<br />

Table 13. Secondary research questions addressed by reviewing existing databases that contain data relating to<br />

surface spills of hydraulic fracturing fluids and wastewater.<br />

Water Cycle Stage<br />

Chemical mixing<br />

Flowback and produced water<br />

Applicable Research Questions<br />

What is currently known about the frequency, severity, and causes of<br />

spills of hydraulic fracturing fluids and additives<br />

What is currently known about the frequency, severity, and causes of<br />

spills of flowback and produced water<br />

3.2.2. Project Introduction<br />

Hydraulic fracturing operations require large quantities of chemical additives, equipment, water,<br />

and vehicles, which may create risks of accidental releases, such as spills or leaks. Surface spills or<br />

releases can occur as a result of events such as tank ruptures, equipment or surface impoundment<br />

failures, overfills, vandalism, accidents, ground fires, or improper operations. Released fluids might<br />

flow into nearby surface water bodies or infiltrate into the soil and near-surface ground water,<br />

potentially reaching drinking water aquifers (NYSDEC, 2011).<br />

Over the past few years, there have been numerous media reports of spills of hydraulic fracturing<br />

fluids and wastewater (US EPA, 2011e). While the media reports have highlighted specific surface<br />

spills of hydraulic fracturing fluids and wastewaters, the frequency and typical causes of these spills<br />

31


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

remain unclear. Additionally, these reports may tend to highlight severe spills and may not<br />

accurately reflect the distribution, number, and severity of spills across the country. The EPA is<br />

compiling information on surface spills of hydraulic fracturing fluids and wastewaters as reported<br />

in federal and state databases to assess the frequency, severity, and causes of spills associated with<br />

hydraulic fracturing. Hydraulic fracturing fluid and wastewater spill information was also collected<br />

from nine hydraulic fracturing service companies and nine oil and gas operators, as discussed in<br />

Sections 3.3 and 3.4, respectively. Together, these data are being used to describe spills of hydraulic<br />

fracturing fluids and wastewater and to identify factors that may lead to potential impacts on<br />

drinking water resources.<br />

3.2.3. Research Approach<br />

There is currently no national repository or database that contains spill data focusing primarily on<br />

hydraulic fracturing operations. In the United States, spills relating to oil and gas operations are<br />

reported to the National Response Center (NRC) and various state regulatory entities. For example,<br />

in Colorado, spills are reported to the Oil and Gas Conservation Commission, within the Department<br />

of Natural Resources, while in Texas, oil and gas related spills are reported to the Texas Railroad<br />

Commission and the Texas Commission on Environmental Quality, depending on which agency has<br />

jurisdiction. The EPA has identified one federal database and databases in five states for review, as<br />

listed in Table 14. The NRC database was selected because it is the only nationwide source of<br />

information on releases of hazardous substances and oil. Spill databases from Colorado, New<br />

Mexico, Pennsylvania, Texas, and Wyoming were chosen for further consideration due to the large<br />

number of hydraulically fractured oil and gas wells found in those states. 23<br />

Table 14. Oil and gas-related spill databases used to compile information on hydraulic fracturing-related incidents.<br />

Source<br />

National Response Center Freedom of<br />

Information Act data<br />

Colorado Oil and Gas Information System<br />

New Mexico Energy, Minerals and Natural<br />

Resources Department<br />

Pennsylvania Department of<br />

Environmental Protection Compliance<br />

Reporting Database<br />

Texas Railroad Commission and Texas<br />

Commission on Environmental Quality<br />

Wyoming Department of Environmental<br />

Quality Water Quality Enforcement Actions<br />

Website<br />

http://www.nrc.uscg.mil/foia.html<br />

http://www.cogcc.state.co.us<br />

https://wwwapps.emnrd.state.nm.us/ocd/ocdpermitting/<br />

Data/Incidents/Spills.aspx<br />

http://www.emnrd.state.nm.us/ocd/Statistics.html<br />

http://www.depreportingservices.state.pa.us/<br />

ReportServer/Pages/ReportViewer.aspx/Oil_Gas/<br />

OG_Compliance<br />

Consolidated Compliance and Enforcement Data System<br />

(not publicly available online)<br />

http://deq.state.wy.us/out/WQenforcementactions.htm<br />

Each of the publicly available databases identified in Table 14 has been searched for spill incidents<br />

related to hydraulic fracturing operations. The search timeframe is limited to incidents between<br />

January 1, 2006, and April 30, 2012, in order to encompass the increase in hydraulic fracturing<br />

23 Based on data provided by nine hydraulic fracturing service companies of oil and gas wells fractured between 2009 and<br />

2010. See Figure 10 in Section 3.3.<br />

32


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

activity seen during that period. To the extent that data are publicly available, electronically<br />

accessible, and readily searchable for spill-related data, the following information is being compiled<br />

about specific hydraulic fracturing-related spill incidents:<br />

• Data source<br />

• Location<br />

• Chemicals/products spilled<br />

• Estimated/reported volume of spill<br />

• Cause of spill<br />

• Reported impact to nearby water resources<br />

• Proximity of the spill to the well or well pad<br />

The information obtained from the NRC and state databases is being reviewed with information<br />

received in response to the EPA’s September 2010 information request to nine hydraulic fracturing<br />

service companies (see Section 3.3) and the EPA’s August 2011 information request to nine oil and<br />

gas operators (Section 3.4). The resulting list of unique spill incidents is being queried to identify<br />

common causes of hydraulic fracturing-related spills, chemicals spilled, the ranges of volumes<br />

spilled, and the potential impacts of these spills to drinking water sources. Because the main focus<br />

of this study is to identify hydraulic fracturing-related spills on the well pad that may impact<br />

drinking water resources, the following topics are not included in the scope of this project:<br />

• Transportation-related spills (except when tanker trucks act as mobile portable storage<br />

containers for chemicals, products, and hydraulic fracturing wastewater used on drilling<br />

sites)<br />

• Drilling mud spills<br />

• Air releases<br />

• Spills associated with disposal through underground injection control wells<br />

• Erosion and sediment control issues<br />

• Spill drills and exercise events (per NRC data)<br />

• Well construction and permitting violations<br />

• Leaks from pipes transporting flowback and produced water from one site to another for<br />

reuse<br />

3.2.4. Status and Preliminary Data<br />

The EPA has initiated work on all publicly available databases listed in Table 14. This section<br />

summarizes the type of information available in each database and lists the criteria being used to<br />

search each database.<br />

National Response Center Freedom of Information Act Data. This database contains nationwide data<br />

on releases of hazardous substances and oil that trigger the federal notification requirements under<br />

33


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

several laws. The NRC is the sole federal point of contact for reporting of all hazardous substances<br />

releases and oil spills. Its information comes from people who arrive on the scene or discover a<br />

spell, then call the NRC hotline or submit a Web-based report form. The information collected by<br />

the NRC during the initial notification call may include the suspected responsible party; the incident<br />

location by county, state, and nearest city; the released material and volume or quantity released;<br />

and a description of the incident, incident causes, affected media, initial known damages, and<br />

remedial actions taken. This information is often based on the estimates made by persons<br />

responding to a spill and may be incomplete. More accurate information may be available once a<br />

response is complete, but this database is not updated with such information.<br />

The data fields that can be used to query the NRC database are listed in Table 15. Many of these<br />

fields only allow searches from a fixed (i.e., drop-down) list, although several of the data fields are<br />

open to any input. None of the search terms in the fixed lists are specific to hydraulic fracturing or<br />

oil and gas exploration and production.<br />

Table 15. Data fields available in the NRC Freedom of Information Act database. ”Fixed list data fields” contain a<br />

fixed list of search terms form which the user can choose. “Open data fields” can receive any input from the user.<br />

Fixed List Data Fields<br />

Type of call<br />

Incident date range<br />

State<br />

County<br />

Incident type<br />

Incident cause<br />

Medium affected<br />

Open Data Fields<br />

NRC report number<br />

Nearest city<br />

Suspected responsible company<br />

Material name<br />

Given the query restrictions, broad searches are being conducted using the listed responsible<br />

company, material name, and incident date range fields (i.e., leaving other fields blank).<br />

The resulting spills are being examined to determine their relevance to this study. Since the<br />

database includes only initial incident reports, information is frequently missing or estimated, such<br />

as total volume spilled. Also, misspellings in the reports or the use of different vocabulary can cause<br />

the search engine to miss relevant incidents.<br />

Colorado. The Colorado Oil and Gas Conservation Commission gathers data regarding pits,<br />

spills/releases, and complaints relating to oil and gas exploration and production. Oil and gas<br />

operators are required to report spills and releases that occur as a result of oil and gas operations,<br />

in accordance with Colorado Oil and Gas Conservation Commission Rule 906 (COGCC, 2011).<br />

Reported information is entered into the Colorado Oil and Gas Information System<br />

Inspection/Incident Database. Each report documents the type of facility, volume spilled and/or<br />

recovered, ground water impacts, depth to shallowest ground water, surface water impacts,<br />

distance to nearest surface water, cause of spill, and a detailed description of the incident. The<br />

34


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

database is searchable by API number, 24 complainant, operator, facility/lease, location, remediation<br />

project number, and document number. Since there is no searchable data field in the database to<br />

indicate whether the spill is related to hydraulic fracturing, the database was queried for all<br />

spill/release reports. Only reports dated from January 1, 2006, to April 30, 2012, were selected for<br />

further review. This search returned over 2,500 reports that are currently being evaluated to<br />

identify incidents related to hydraulic fracturing activities.<br />

New Mexico. The Oil Conservation Division of the State of New Mexico Energy, Minerals and Natural<br />

Resources Department tracks information, in two separate databases, on both spill incidents and<br />

incidents where liquids in pits have contaminated ground water. Release Notification and<br />

Corrective Action forms are submitted to the Oil Conservation Divisions District offices. Spills can<br />

be reported by industry representatives or state agency personnel.<br />

The spills database is searchable by facility and well names, incident type, operator, location, lease<br />

type, spilled material, spill cause, spill source, and the spill referrer (person who reported the<br />

incident). The database was initially searched using the spill material, spill cause, and spill source<br />

data fields. Each of these fields can only be searched using the preset search terms listed in Table<br />

16. The initial search was conducted using the search terms in bold in Table 16. The EPA is<br />

currently examining the resulting list of spills to determine their relevancy to this study and is<br />

considering running additional queries to collect more information.<br />

24 The API (American Petroleum Institute) number is a unique, permanent, numeric identifier assigned to each well<br />

drilled for oil and gas in the United States.<br />

35


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table 16. Preset search terms available for the spill material, spill cause, and spill source data fields in the New<br />

Mexico Oil Conservation Division Spills Database. Terms in bold have been searched.<br />

Spill Material Spill Cause Spill Source<br />

All All All<br />

Acid Blowout Coupling<br />

Brine water Corrosion Gas compression station<br />

B.S. & W (basic sediment & water) Equipment failure Dump line<br />

Chemical (specify) Fire Motor<br />

Condensate Freeze Flowline—injection<br />

Diesel Human error Flowline—production<br />

Drilling mud/fluid Lightning Frac tank<br />

Glycol Other Fitting<br />

Gasoline Normal operations Injection header<br />

Gelled brine (frac fluid) Vandalism Other (specify)<br />

Hydrogen sulfate Vehicular accident Pit (specify)<br />

Crude oil<br />

Motor oil<br />

Natural gas (methane)<br />

Natural gas liquids<br />

Lube oil<br />

Other (specify)<br />

Produced water<br />

Unknown<br />

Pipeline (any)<br />

Production tank<br />

Pump<br />

Separator<br />

Transport<br />

Unknown<br />

Valve<br />

Well<br />

Water tank<br />

The database containing information regarding contamination of ground water due to pits tracks<br />

only the current company, facility name, tracking number, county, location, and status of the<br />

contamination incidents. Details regarding the contamination incident and the relation of the event<br />

to hydraulic fracturing are not included. Additional research is needed to determine if the pit<br />

information is related to hydraulic fracturing.<br />

Pennsylvania. The Pennsylvania Department of Environmental Protection’s Compliance Reporting<br />

Database provides information on oil and gas inspections, violations, enforcement actions, and<br />

penalties assessed and collected. Users can search the database according to the following fixedvariable<br />

data fields: county, municipality, date inspected, operator, Marcellus only, 25 inspections<br />

with violations only, and resolved violations only.<br />

Table 17 displays the total number of incidents retrieved for four different queries, all using a date<br />

range of January 1, 2006, to April 30, 2012.<br />

25 This data field was recently changed to “unconventional only” (last accessed July 6, 2012).<br />

36


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table 17. Total number of incidents retrieved from the Pennsylvania Department of Environmental Protection's<br />

Compliance Reporting Database by varying inputs in the “Marcellus only” and inspections with “violations only data<br />

fields.” In all cases, “no” was entered in the “resolved violations only” field.<br />

Marcellus Only<br />

Inspections with<br />

Violations Only<br />

Yes<br />

No<br />

Yes<br />

Yes<br />

No<br />

Yes<br />

No<br />

No<br />

* Error message received when formatting results of this query.<br />

Total Number of<br />

Incidents Retrieved<br />

25,687<br />

4,319<br />

18,700<br />

Unknown*<br />

The queries shown in Table 17 returned information collected during inspections that found<br />

violations and/or when spills are reported. An incident or inspection may have multiple violations,<br />

leading to a large total number of violations retrieved from the database. The EPA’s initial effort<br />

focused on the query that returned the fewest violations, which totaled 4,319 inspections with<br />

violations specific to the Marcellus Shale region. Inspection and violation comment fields for each<br />

incident are being reviewed to identify incidents related to hydraulic fracturing activities.<br />

Texas. Representatives of the Railroad Commission of Texas, the Texas Commission on<br />

Environmental Quality, and the Texas General Land Office have confirmed that there is no central<br />

database in Texas on hydraulic fracturing-related spills. In Texas, a memorandum of understanding<br />

between the Railroad Commission and Commission on Environmental Quality identifies the<br />

jurisdiction of these agencies over waste materials resulting from exploring, developing, producing,<br />

and refining oil and gas. Pursuant to this understanding, oil and gas operators are required to<br />

report spills to the Railroad Commission, which maintains a publicly available database of spills of<br />

petroleum, oil, and condensate. The EPA has reviewed this database and determined that it does<br />

not include chemical spills; most of the spills reported in the database are crude oil spills.<br />

Therefore, there will be no further analysis of this database.<br />

The Commission on Environmental Quality is Texas’ lead agency in responding to spills of all<br />

hazardous substances that may cause pollution or lower air quality pursuant to the Texas<br />

Hazardous Substances Spill Prevention and Control Act (Texas Water Code §26.261). The<br />

Commission on Environmental Quality may generate an investigation, inspection, or complaint<br />

report in response to emergency spill notifications. These reports are submitted to the state’s<br />

Consolidated Compliance and Enforcement Data System. However, the investigation and inspection<br />

reports in this database are not available electronically on the Texas Commission on Environmental<br />

Quality’s website or at their Central Files Room.<br />

Other attempts were made to obtain information on potential ground water contamination<br />

incidents related to hydraulic fracturing by examining the Joint Groundwater Monitoring and<br />

Contamination Reports prepared by the Texas Groundwater Protection Committee; this effort was<br />

unsuccessful in getting the relevant incident details. The abovementioned searches for hydraulic<br />

fracturing spill-related data may not be an exhaustive investigation of all available information<br />

from Texas’ state agencies or organizations, but other publicly available sources of information<br />

have not been located at this time.<br />

37


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Wyoming. The Wyoming Department of Environmental Quality maintains a publicly available<br />

database of water quality enforcement actions. This database includes reports of water quality<br />

violations categorized by the year they occurred, from 2006 to 2012. None of the reports<br />

differentiate between hydraulic fracturing-related incidents and those due to other stages of oil and<br />

gas development. Many of the oil and gas-related violations were for CBM produced water<br />

discharges, such as to surface water. Due to the lack of information to differentiate between<br />

hydraulic fracturing-related incidents and other oil and gas-related incidents, there will be no<br />

further analysis of this dataset.<br />

The spills database analysis has several important limitations:<br />

• Potential underreporting. This affects the EPA’s ability to assess the number or frequency<br />

of hydraulic fracturing-related spill incidents, since it is likely that some spills are not<br />

reported to the NRC or state agencies.<br />

• Variation in reporting requirements for different sources. This makes it difficult to<br />

categorize reported spills as hydraulic fracturing-related and to comprehensively identify<br />

the causes, chemical identity, and volumes of hydraulic fracturing-related spills.<br />

• The lack of electronic accessibility of some state-reported data on oil and gas-related spills<br />

and emergency responses. This also significantly impacts the comprehensiveness of the<br />

available information.<br />

3.2.5. Next Steps<br />

As noted, the EPA is reviewing the list of spill incidents generated by searching the NRC, Colorado,<br />

New Mexico, and Pennsylvania databases to identify incidents related to hydraulic fracturing<br />

activities. Spill incidents identified through this review will be combined with data received from<br />

nine hydraulic fracturing service companies (see Section 3.3) and nine oil and gas operators<br />

(Section 3.4) to create a master database of hydraulic fracturing-related spills from these sources.<br />

The compiled information will be examined to identify, where possible, common causes of<br />

hydraulic fracturing-related spills, chemicals spilled, and ranges of volumes spilled. Specific steps<br />

will then include:<br />

• Creating a reference table of information gathered from all incidences determined to be<br />

related to hydraulic fracturing.<br />

• Reviewing this reference table for trends in the causes and volumes of hydraulic<br />

fracturing-related spills.<br />

3.2.6. Quality Assurance Summary<br />

The QAPP for the analysis of publicly available information on surface spills related to hydraulic<br />

fracturing, “Hydraulic Fracturing (HF) Surface Spills Data Analysis (Version 1),” was approved on<br />

August 6, 2012 (US EPA, 2012l). The project underwent a technical systems audit (TSA) by the<br />

designated EPA QA Manager on August 27, 2012. The methods and products being developed under<br />

the project adhered to the approved QAPP, and no corrective actions were identified.<br />

38


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

3.3. Service Company Analysis<br />

3.3.1. Relationship to the Study<br />

The EPA asked nine hydraulic fracturing service companies for information about hydraulic<br />

fracturing operations conducted from 2005 to 2010. The data are being analyzed for information<br />

that can be used to inform answers to the research questions in Table 18.<br />

Table 18. Secondary research questions addressed by analyzing data received from nine hydraulic fracturing service<br />

companies.<br />

Water Cycle Stage<br />

Water acquisition<br />

Chemical mixing<br />

Well injection<br />

Flowback and<br />

produced water<br />

Applicable Research Questions<br />

• How much water is used in hydraulic fracturing operations, and what are<br />

the sources of this water<br />

• What is currently known about the frequency, severity, and causes of<br />

spills of hydraulic fracturing fluids and additives<br />

• What are the identities and volumes of chemicals used in hydraulic<br />

fracturing fluids, and how might this composition vary at a given site and<br />

across the country<br />

• How effective are current well construction practices at containing gases<br />

and fluids before, during, and after fracturing<br />

• Can subsurface migration of fluids or gases to drinking water resources<br />

occur and what local geologic or man-made features may allow this<br />

• How might hydraulic fracturing fluids change the fate and transport of<br />

substances in the subsurface through geochemical interactions<br />

• What is currently known about the frequency, severity, and causes of<br />

spills of flowback and produced water<br />

• What is the composition of hydraulic fracturing wastewaters, and what<br />

factors might influence this composition<br />

3.3.2. Project Introduction<br />

Hydraulic fracturing is typically performed by a service company under a contract with the oil or<br />

gas production well operator. The service companies possess detailed information regarding the<br />

implementation of hydraulic fracturing, from design through fracturing. In September 2010, the<br />

EPA requested information from nine companies on the chemical composition of hydraulic<br />

fracturing fluids used from 2005 to 2010, standard operating procedures (SOPs), impacts of<br />

chemicals on human health and the environment, and the locations of oil and gas wells<br />

hydraulically fractured in 2009 and 2010. The EPA is analyzing the information received from the<br />

service companies to better understand current hydraulic fracturing operating practices and to<br />

answer the research questions listed above.<br />

Service Companies Selected. Nine service companies received the information request: BJ Services<br />

Company, Complete Production Services, Halliburton, Key Energy Services, Patterson-UTI Energy,<br />

RPC, Schlumberger, Superior Well Services, and Weatherford International. These companies<br />

reflect a range of industry market share and variation in company size. The EPA estimated that BJ<br />

Services Company, Halliburton, and Schlumberger performed approximately 95% of hydraulic<br />

fracturing services in the United States in 2003 (US EPA, 2004b), and the three companies reported<br />

39


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

the highest annual revenues for 2009 of the nine companies selected for the information request. 26<br />

The remaining six companies represent mid-sized and small companies performing hydraulic<br />

fracturing services between 2005 and 2009. 27 Table 19 shows the annual revenue, number of<br />

employees, and company services reported by the companies to the US Securities and Exchange<br />

Commission in the 2009 Form 10-K.<br />

Table 19. Annual revenue and approximate number of employees for the nine service companies selected to receive<br />

the EPA’s September 2010 information request. The companies reflect a range of industry market share and<br />

company sizes. Information was obtained from Form 10-K, filed with the US Securities and Exchange Commission in<br />

2009.<br />

Company<br />

Annual Revenue for<br />

2009 (Millions)<br />

Number of<br />

Employees<br />

(Approximate)<br />

BJ Services Company* $4,122 14,400<br />

Complete Production Services $1,056 5,200<br />

Halliburton $14,675 51,000<br />

Key Energy Services $1,079 8,100<br />

Patterson-UTI Energy $782 4,200<br />

RPC $588 2,000<br />

Schlumberger $22,702 77,000<br />

Superior Well Services $399 1,400<br />

Weatherford International $8,827 52,000<br />

* BJ Services reports on a fiscal year calendar ending on September 30.<br />

Three of the nine service companies that reported information to the EPA were acquired by other<br />

companies since 2010. Baker Hughes completed the purchase of BJ Services Company in April 2010,<br />

Patterson-UTI Energy purchased Key Energy Services in October 2010, and Superior Well Services<br />

acquired Complete Production Services in February 2012.<br />

3.3.3. Research Approach<br />

The EPA received responses to the September 2010 information request from each of the nine<br />

service companies. Data and information relevant to the research questions posed above were<br />

collected and organized in Microsoft Excel spreadsheets and Microsoft Access databases. Each<br />

company reported information in various organizational formats and using different descriptive<br />

terms; therefore, the EPA has put all nine datasets into a consistent format for analysis and<br />

resolving any issues associated with terminology, data gaps, or inconsistencies. This selection of<br />

information serves as the basis for targeted queries and data summaries described below. The<br />

queries and data summaries have been designed to answer the secondary research questions listed<br />

in Table 18.<br />

Much of the data and information received by the EPA was claimed to be confidential business<br />

information (CBI) under the Toxic Substances Control Act (TSCA). Five of the nine companies,<br />

26 Information was obtained from the 2009 Form 10-K, filed with the US Securities and Exchange Commission.<br />

27 Annual revenue and number of employees were used as indicators of company size.<br />

40


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

however, also provided non-confidential information. 28 Because the majority of the information has<br />

been claimed as CBI, the analyses described below are being conducted in accordance with the<br />

procedures outlined in the EPA’s TSCA CBI Protection Manual (US EPA, 2003b). All results are<br />

treated as CBI until determinations are made or until masking has been done to prevent disclosure<br />

of CBI information.<br />

Summary of Service Company Operations. The EPA is using information provided by the companies<br />

to write a narrative description of the range of their operations, which includes information on the<br />

role of the service companies in each stage of the hydraulic fracturing water cycle.<br />

Information has been compiled on the number and location of wells hydraulically fractured by the<br />

nine service companies between September 2009 and October 2010, resulting in a map that<br />

displays the number of wells fractured per county as reported by the companies. This information<br />

is intended to illustrate the intensity and geographic distribution of hydraulic fracturing activities<br />

by these companies.<br />

Water Acquisition. The following information from the service company data on volumes, quality,<br />

and sources of water used in hydraulic fracturing fluids is being summarized and will include:<br />

• Water use by shale play. The range of water volumes used based on the shale play in which<br />

the well is located. (The companies did not provide information on geologic formations<br />

other than shale.)<br />

• Procedures and considerations relating to water acquisition. Summary of any SOPs, water<br />

quality requirements, water source preferences, and decision processes described in the<br />

submissions from the nine service companies.<br />

Chemical Mixing. The following information collected from the service companies is being<br />

assembled to identify the composition of different hydraulic fracturing fluid formulations and the<br />

factors that influence formulation composition:<br />

• Chemical name<br />

• Chemical formula<br />

• Chemical Abstracts Service Registration Number (CASRN)<br />

• Material Safety Data Sheets (MSDSs) for each fluid product<br />

• Concentration of each chemical in each fluid product<br />

• Manufacturer of each product and chemical<br />

• Purpose and use of each chemical in each fluid product<br />

28 The non-confidential information is available on the federal under docket number EPA-HQ-ORD-2010-0674 or via<br />

http://www.regulations.gov/#!searchResults;rpp=10;po=0;s=epa-hq-ord-2010-0674.<br />

41


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

For the purposes of the analysis, the EPA defines a “product” as an additive composed of a single<br />

chemical or several chemicals. A “chemical” is an individual chemical included in a product. A “fluid<br />

formulation” is the entire suite of products and carrier fluid injected into a well during hydraulic<br />

fracturing. The following information from the service company data on chemicals, products, and<br />

fluid formulations is being summarized:<br />

• Formulations, products, and product function. The formulations reported by the nine<br />

service companies and the number and types of products used in those formulations.<br />

• Products, chemicals in those products and concentrations, and manufacturer of each<br />

product. The chemicals used in each product may be used in conjunction with the<br />

formulations data (described in the previous bullet) to discern the chemicals used in each<br />

formulation. The manufacturer of each product will also be included.<br />

• Number of products reported for a given product function and the frequency with which a<br />

product function is reported in the formulations data. The product function with the<br />

greatest number of products and the product function that is most often used in<br />

formulations.<br />

• Number of products and chemicals for each type of formulation. The chemicals and products<br />

for various types of formulations and a description of the average number of products and<br />

chemicals for each formulation type, as well as the sample size for each population and<br />

common product functions for each formulation type.<br />

• Typical loadings for each group of products of a given product function and for each fluid<br />

formulation type. The typical proportion of a product in a formulation. Typical loading<br />

values (e.g., gallons per thousand gallons) indicate an amount or volume of a product<br />

added to a volume of fracturing fluids rather than an accurate representation of the<br />

concentration of a particular product or the chemical constituents of a product in a fluid<br />

formulation.<br />

Information provided by the companies relating to surface spills of hydraulic fracturing fluids and<br />

chemicals has been compiled, resulting in a table of specific spill incidences. The table includes<br />

information on the location, composition, volume, cause, and any reported impacts of each spill.<br />

This information will be used in the larger analysis of surface spills reported in federal and state<br />

databases (Section 3.2).<br />

Well Injection. The EPA requested information regarding the hydraulic fracturing service<br />

companies’ procedures for establishing well integrity, procedures used during well injections, and<br />

response plans to address unexpected circumstances (e.g., unexpected pressure changes during<br />

injection). Information provided by the companies will be used to write a narrative description of<br />

the range of operations conducted by this sample of service companies.<br />

Flowback and Produced Water. Although this information was not requested, the EPA received<br />

some documents and information that referenced flowback and produced water, including policies,<br />

practices, and procedures employed by companies to determine estimated volumes and<br />

management options. The EPA has reviewed this information as well as information relevant to the<br />

42


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

frequency, severity, and causes of flowback and produced water spills and the composition of<br />

hydraulic fracturing wastewaters. The outputs of the analysis will include the following:<br />

• Reported spills of flowback and produced water. Information on the composition of the fluid<br />

spilled, the volume spilled, the reported cause of the spill, and any reported impacts to<br />

nearby water resources. This information will be integrated into the larger analysis of<br />

surface spills reported in federal and state databases (see Section 3.2).<br />

• Reported compositions of hydraulic fracturing wastewater. Information on the chemical and<br />

physical properties of hydraulic fracturing wastewater, such as the identities of analytes of<br />

interest and reported concentration ranges. To the extent possible, this information will be<br />

organized according to geologic and geographic location as well as time after fluid<br />

injection.<br />

• Flowback and produced water management. Where possible, information about the role of<br />

hydraulic fracturing service companies in handling flowback and produced water will be<br />

described.<br />

3.3.4. Status and Preliminary Data<br />

Preliminary data analyses of service company operations, water acquisition, chemical mixing, and<br />

flowback and produced water has been completed and the analysis of well injection information<br />

has begun. The EPA has met with representatives from each of the nine hydraulic fracturing service<br />

companies to discuss their responses to the September 2010 information request. Information<br />

gathered during these meetings has been used to inform the data analysis and to ensure that<br />

confidential information is protected. As of September 2012, the EPA continues to clarify the<br />

information reported and to work with the nine hydraulic fracturing service companies to release<br />

information originally designated as CBI without compromising trade secrets.<br />

Service Company Operations. As a group, the nine service companies reported that they<br />

hydraulically fractured 24,925 wells in the United States in 2009 and 2010. The companies<br />

reported the number of wells per county, which is displayed for all companies in Figure 10.<br />

43


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Figure 10. Locations of oil and gas production wells hydraulically fractured between September 2009 and October<br />

2010. The information request to service companies (September 2010) resulted in county-scale locations for 24,925<br />

wells. The service company wells represented in this map include only 24,879 wells because the EPA did not receive<br />

locational information for 46 of the 24,925 reported wells. (ESRI, 2010a, b; US EPA, 2011a)<br />

Chemical Mixing. The service companies reported a total of 114 example formulations and 1,858<br />

unique producets, which consist of 677 unique chemicals, used by the service companies between<br />

September 2005 and 2010. 29 Table 20 shows the number of formulations, products, and chemicals<br />

reported by each of the nine service companies; the totals for products and chemical constituents in<br />

Table 20 reflect use by multiple companies and are therefore greater than the sum of unique<br />

products and chemical constituents. The formulations reported to the EPA are not comprehensive,<br />

as each service company chose them as examples of the fluids they use.<br />

29 Products and chemical constituents noted here are unique and may have been reported multiple times by the service<br />

companies.<br />

44


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table 20. Formulations, products, and chemicals reported as used or distributed by the nine service companies<br />

between September 2005 and September 2010.<br />

Company Formulations Products* Chemical Constituents †<br />

BJ Services 37 401 118<br />

Key Energy Services 16 180 119<br />

Halliburton 15 450 304<br />

RPC 13 182 128<br />

Schlumberger 11 110 61<br />

Patterson-UTI Energy 10 67 67<br />

Weatherford International 6 214 180<br />

Complete Production Services 3 122 92<br />

Superior Well Services 3 312 117<br />

* Companies reported examples of formulations, which did not contain all of the products reported to the EPA.<br />

†<br />

Not all products have reported chemicals.<br />

Non-confidential hydraulic fracturing chemicals reported by the companies appear in Appendix A,<br />

along with chemicals reported from publicly available sources.<br />

Well Injection. Seven service companies reported 231 protocols to the EPA. The protocols describe<br />

the procedures used by the companies for many aspects of field and laboratory work, including site<br />

and infrastructure planning, chemical mixing and design of fracturing fluid formulations, health and<br />

safety practices, well construction, and hydraulic fracturing. The EPA is analyzing the information<br />

to assess how hydraulic fracturing service companies use SOPs, to better understand how well<br />

integrity is established prior to fracturing, and to evaluate procedures used during well injection.<br />

Flowback and Produced Water. Data provided by the companies indicate that the company<br />

conducting the fracturing is often not the same company that manages the flowback process. Five of<br />

the companies responded that they do not provide flowback services, although one of these<br />

companies provides analytical support to operators for the testing of flowback water for potential<br />

reuse. Two of the nine stated that they provide flowback services independent of their hydraulic<br />

fracturing services. For another two companies, the EPA received no information clearly describing<br />

role regarding flowback services. Only one company provided detailed information on flowback<br />

management.<br />

3.3.5. Next Steps<br />

All analyses will undergo a QA review before being compiled in a final report. The EPA will continue<br />

to work with each of the nine companies to determine how best to summarize the results so that<br />

CBI is protected while providing information in a transparent manner.<br />

3.3.6. Quality Assurance Summary<br />

The QAPP for the analysis of data received from nine service companies, “Analysis of Data Received<br />

from Nine Hydraulic Fracturing (HF) Service Companies (Version 1),” was approved on August 1,<br />

2012 (US EPA, 2012h). A TSA on the work was conducted by designated EPA QA Manager on<br />

August 28, 2012, to review the methods being used and work products being developed with the<br />

data. The work accurately reflected what is described in the QAPP, and no corrective actions were<br />

45


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

identified. In addition, the EPA’s contractor, Eastern Research Group, has been involved with<br />

collecting and compiling data submitted from the nine hydraulic fracturing service companies.<br />

Eastern Research Group’s QAPP was approved on January 19, 2011 (Eastern Research Group Inc.,<br />

2011).<br />

3.4. Well File Review<br />

3.4.1. Relationship to the Study<br />

The well file review provides an opportunity to assess well construction and hydraulic fracturing<br />

operations, as reported by the companies that own and operate oil and gas production wells.<br />

Results from the review will inform answers to the secondary research questions listed in Table 21.<br />

Table 21. Secondary research questions addressed by the well file review research project.<br />

Water Cycle Stage<br />

Water acquisition<br />

Chemical mixing<br />

Well injection<br />

Flowback and produced water<br />

Wastewater treatment and<br />

waste disposal<br />

Applicable Research Questions<br />

• How much water is used in hydraulic fracturing operations, and<br />

what are the sources of this water<br />

• What is currently known about the frequency, severity, and<br />

causes of spills of hydraulic fracturing fluids and additives<br />

• What are the identities and volumes of chemicals used in<br />

hydraulic fracturing fluids, and how might this composition vary<br />

at a given site and across the country<br />

• If spills occur, how might hydraulic fracturing chemical additives<br />

contaminate drinking water resources<br />

• How effective are current well construction practices at<br />

containing gases and fluids before, during, and after fracturing<br />

• Can subsurface migration of fluids and gases to drinking water<br />

resources occur and what local geologic or man-made features<br />

may allow this<br />

• What is currently known about the frequency, severity, and<br />

causes of spills of flowback and produced water<br />

• What is the composition of hydraulic fracturing wastewaters,<br />

and what factors might influence this composition<br />

• If spills occur, how might hydraulic fracturing wastewater<br />

contaminate drinking water resources<br />

• What are the common treatment and disposal methods for<br />

hydraulic fracturing wastewaters, and where are these methods<br />

practiced<br />

3.4.2. Project Introduction<br />

The process of planning, designing, permitting, drilling, completing, and operating oil and gas wells<br />

involves many steps, all of which are ultimately controlled by the company that owns or operates<br />

the well, referred to as the “operator.” Assisting the operator are service companies that provide<br />

specialty services, such as seismic surveys, lease acquisition, road and pad building, well drilling,<br />

logging, cementing, hydraulic fracturing, water and waste hauling, and disposal. Some operators<br />

can perform some of these services on their own and some rely exclusively on service companies.<br />

During the development and production of oil and gas wells, operators receive documentation from<br />

service companies about site preparation and characteristics, well design and construction,<br />

hydraulic fracturing, oil and gas production, and waste management. Operators typically maintain<br />

46


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

much of this information in an organized file, which cumulatively represents the history of the well.<br />

The EPA refers to this file as a “well file.” Some of the information in a well file may be required by<br />

law to be reported to state oil and gas agencies, and some of the information may be considered CBI<br />

by the operator.<br />

For this project, the EPA is scrutinizing actual well files from hydraulic fracturing operations in<br />

different geographic areas that are operated by companies of various sizes. These wells include<br />

vertical, horizontal, and deviated wells that produce oil, gas, or both from differing geological<br />

environments. This review is providing information that can be used to identify practices that may<br />

impact drinking water resources.<br />

3.4.3. Research Approach<br />

While a portion of the data needed for this project is reported to state oil and gas agencies, the<br />

complete dataset is available only in the well files compiled by oil and gas operators. 30 Further,<br />

different states have different reporting requirements. As a result, the EPA selected 350 well<br />

identifiers believed to represent oil and gas production wells hydraulically fractured by the nine<br />

hydraulic fracturing service companies and requested the corresponding well files from operators<br />

associated with those wells. 31 This section describes the process used by the EPA to select well files<br />

for review, the information requested, and the planned analyses.<br />

Well File Selection. The EPA used a list of hydraulically fractured oil and gas wells provided to the<br />

agency by the nine hydraulic fracturing service companies (referred to hereafter as the “service<br />

company well list”) to select 350 specific well identifiers associated with nine oil and gas<br />

operators. 32 The service company well list obtained by the EPA contains 24,925 well identifiers<br />

associated with wells that were reported to have been hydraulically fractured between September<br />

2009 and October 2010 (Figure 10) and identifies 1,146 oil and gas operators. This compiled list<br />

includes, for each well, a well identifier, the operator’s name, and the well’s state and county<br />

location.<br />

Counties containing the 24,925 well identifiers were grouped into four geographic regions<br />

according to a May 9, 2011, map of current and prospective shale gas plays within the lower 48<br />

states (US EIA, 2011c). 33 If any portion of a county was within one of the shale gas plays defined on<br />

the map, the entire county was assigned to that shale play and the corresponding geographic<br />

region. The four regions—East, South, West, and Other—are shown in Figure 11 with the<br />

corresponding number of wells in each region. Counties outside the shale gas plays were grouped<br />

30 The EPA analyzed several state oil and gas agency websites and estimated that it would find less than 15% of the<br />

necessary data from websites to answer the research questions.<br />

31 Oil and gas production wells are generally assigned API numbers by state oil and gas agencies, a unique 10-digit<br />

number. Wells may also be commonly identified by a well name that is designated by the operator. The EPA considers<br />

both of these to be well identifiers.<br />

32 The EPA used the service company well list because it is unaware of the existence of a single list showing all oil and gas<br />

production wells in the United States, their operators, and whether each well has been hydraulically fractured.<br />

33 Wells within a designated shale play on the map are not guaranteed to be producing from that shale; they could be<br />

producing from rock formations within the same stratigraphic column.<br />

47


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

into the Other region, which includes areas where oil and gas is produced from a variety of rock<br />

formations. 34 This grouping process allowed the EPA to select wells that reflect the geographic<br />

distribution of hydraulically fractured oil and gas wells.<br />

A list of operators and their corresponding total well count was sorted by well count from highest<br />

to lowest. Operators with fewer than 10 well identifiers were removed, resulting in a final list of<br />

266 operators and 22,573 wells. The resulting operators were categorized as “large,” “medium,” or<br />

“small.” Large operators were defined as those that accounted for the top 50% of the well<br />

identifiers on the list, medium operators for the next 25% and small operators for the last 25%. As<br />

a result, there were 17 large operators, 86 medium operators, and 163 small operators. To ensure<br />

that the final selected well identifiers would have geographic diversity among large operators, each<br />

large operator was assigned to one geographic region that contained a large number of its well<br />

identifiers.<br />

One large operator was randomly chosen from each of the regions (i.e., one large operator from<br />

each of the East, South, West, and Other regions), for a total of four large operators. Two medium<br />

operators and three small operators were also chosen, with no preference for geographic region.<br />

This resulted in the selection of nine operators: Clayton Williams Energy, ConocoPhillips, EQT<br />

Production, Hogback Exploration, Laramie Energy, MDS Energy, Noble Energy, SandRidge Energy,<br />

and Williams Production.<br />

34 Forty-six well identifiers had unknown counties and have been included in the Other region for the purposes of this<br />

analysis.<br />

48


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Figure 11. Locations of oil and gas production wells hydraulically fractured from September 2009 through October<br />

2010. The information request to service companies (September 2010) resulted in county-scale locations for 24,925<br />

wells. The service company wells are represented above as regional well summaries and summarize only 24,879<br />

wells because the EPA did not have locational information for 46 of the 24,925 reported wells. (ESRI, 2010a, b; US<br />

EPA, 2011a)<br />

The nine operators were associated with 2,455 well identifiers. The EPA initially chose 400 of those<br />

2,455 well identifiers to request the associated well files for its analysis. The selection of 400 well<br />

identifiers required balancing goals of maximizing the geographic diversity of wells and maximizing<br />

the precision of any forthcoming statistical estimates. The well identifiers were chosen using an<br />

optimization algorithm that evaluated the statistical precision given different allocations across<br />

operating company/shale play combinations. The algorithm identified a solution given four<br />

constraints:<br />

49


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

• Select all well identifiers for the three small operators whose total number of well<br />

identifiers was fewer than 35. For all other operators, keep the number of selected well<br />

identifiers between 35 and 77.<br />

• Have at least two well identifiers (or one if there is only one) from each combination of a<br />

large operator and geographic region.<br />

• Keep the regional distribution of sampled well identifiers close to the regional distribution<br />

of all 24,925 well identifiers on the initial service company well list.<br />

• Keep the expected sampling variance due to unequal weights relatively small.<br />

Due to resource and time constraints, the EPA subsequently decided to review 350 well files, so 50<br />

of the 400 selected well identifiers were randomly removed. This sample size is large enough to be<br />

considered reasonably representative of the total number of wells hydraulically fractured by the<br />

nine service companies in the United States during the specified time period.<br />

Data Requested. An information request letter was sent in August 2011 to the nine operators<br />

identified above, asking for 24 distinct items organized into five topic areas: (1) geologic maps and<br />

cross sections; (2) drilling and completion information; (3) water quality, volume, and disposition;<br />

(4) hydraulic fracturing; and (5) environmental releases. 35 Table 22 shows the potential<br />

relationship between the five topic areas and the stages of the hydraulic fracturing water cycle.<br />

Table 22. The potential relationship between the topic areas in the information request and the stages of the<br />

hydraulic fracturing water cycle.<br />

Information Request Topic Areas<br />

Water Cycle Geologic Maps Drilling and Water Quality,<br />

Stage<br />

Hydraulic<br />

and Cross - Completion Volume, and<br />

Fracturing<br />

Sections Information Disposition<br />

Water acquisition <br />

Environmental<br />

Releases<br />

Chemical mixing <br />

Well injection <br />

Flowback and<br />

produced water<br />

Wastewater<br />

treatment and<br />

waste disposal<br />

<br />

Well File Review and Analysis. The EPA received responses to the August 2011 information request<br />

from each of the nine operators. Data and information contained in the well files is being extracted<br />

from individual well files and compiled in a single Microsoft Access database. All data in the<br />

database are linked by the well’s API number; this process is described in more detail in the QAPP<br />

for this research project (US EPA, 2012j).<br />

<br />

<br />

35 See the text of the information request for the specific items requested under each topic area. The information request<br />

can be found at http://www.epa.gov/<strong>hf</strong>study/August_2011_request_letter.pdf.<br />

50


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Information in the database is being used to design queries that will inform answers to the research<br />

questions listed in Table 21. Examples of queries being designed include:<br />

• What sources and volumes of water are used for hydraulic fracturing fluids<br />

• How many well files contain reports of chemicals spilled during hydraulic fracturing, and<br />

do the reports show whether the spills led to any impacts to drinking water resources<br />

• How many wells have poor cement bonds immediately above the uppermost depth being<br />

hydraulically fractured This may indicate that the cement sheath designed to isolate the<br />

target zone being stimulated may fail, potentially leading to gas and fluid migration up the<br />

wellbore.<br />

• How many well files contain reports of flowback or produced water spilled, and do the<br />

reports show whether the spills lead to any impacts to drinking water resources<br />

• What are the reported treatment and/or disposal methods for the wastewater generated<br />

from hydraulic fracturing<br />

3.4.4. Status and Preliminary Data<br />

Of the 350 well identifiers selected for analysis, the EPA received information on 334 wells. One of<br />

these was never drilled, ultimately providing the EPA with well files for 333 drilled wells. 36 Table<br />

23 lists the number of wells for which valid data were provided by each operator and their<br />

designated company size.<br />

Table 23. Number of wells for which data were provided by each operator. Company size, as determined for this<br />

analysis, is also listed. The nine operators provided data on a total of 333 oil and gas production wells.<br />

Operator Company Size Number of Wells<br />

Noble Energy Large 67<br />

ConocoPhillips Large 57<br />

Williams Production Large 50<br />

Clayton Williams Energy Medium 36<br />

SandRidge Energy Medium 35<br />

EQT Production Large 29<br />

MDS Energy Small 24<br />

Laramie Energy Small 21<br />

Hogback Exploration Small 14<br />

Total 333<br />

Figure 12 shows a map of the 333 well locations. The well locations are distributed within 13<br />

states: Arkansas, Colorado, Kentucky, Louisiana, New Mexico, North Dakota, Oklahoma,<br />

Pennsylvania, Texas, Utah, Virginia, West Virginia, and Wyoming.<br />

36 Sixteen of the 350 well identification numbers were not valid for this project: 13 were duplicate entries, one was in<br />

Canada, one was not a well, and one was not actually owned by the selected operator. In total, roughly 5% of the 350 well<br />

identifiers chosen for review by the EPA do not correspond to oil and gas wells that have been hydraulically fractured.<br />

This provides a rough assessment of the accuracy of the original data received from the nine hydraulic fracturing service<br />

companies (the service company well list).<br />

51


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Figure 12. Locations of 333 wells (black points) selected for the well file review. Also shown are the locations of oil<br />

and gas production wells hydraulically fractured from September 2009 through October 2010. The information<br />

request to service companies (September 2010) resulted in county-scale locations for 24,925 wells. The service<br />

company wells are represented above as regional well summaries and summarize only 24,879 wells because the<br />

EPA did not have locational information for 46 of the 24,925 reported wells. (ESRI, 2010a, b; US EPA, 2011a, d)<br />

The EPA received approximately 9,670 electronic files in response to the August 2011 information<br />

request. The amount of information received varied from one well file to another. Some well files<br />

included nearly<br />

all<br />

of the<br />

information requested, while others were missing<br />

information<br />

on<br />

entire<br />

topical areas.<br />

Some of the<br />

data received<br />

were<br />

claimed<br />

as CBI<br />

under<br />

TSCA.<br />

The EPA<br />

has<br />

contacted<br />

all<br />

nine of the oil and gas operators<br />

to clarify<br />

its understanding<br />

of the data, where<br />

necessary, and to<br />

discuss<br />

how to<br />

depict<br />

the<br />

well<br />

file<br />

data while still protecting confidential information.<br />

The analyses<br />

described<br />

in the previous<br />

section<br />

are<br />

being<br />

performed<br />

according<br />

to<br />

CBI procedure<br />

s (US<br />

EPA,<br />

2003b)<br />

, and<br />

the<br />

results<br />

are considered<br />

CBI<br />

until<br />

determinations<br />

are<br />

made<br />

or until data<br />

masking has<br />

been done to<br />

prevent<br />

release of CBI<br />

information.<br />

The EPA is extracting available data from the well files that can be used to answer research<br />

questions related to all stages of the hydraulic fracturing water cycle. As of September 2012, the<br />

52


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

EPA had extracted, and continues to extract, the following available information from all of the well<br />

files:<br />

• Open-hole log analysis of lithology, hydrocarbon shows, and water salinity<br />

• Chemical analyses of various water samples<br />

• Well construction data<br />

• Cement reports<br />

• Cased-hole logs, including identifying cement tops and bond quality<br />

Other data to be extracted includes the following:<br />

• Source of water used for hydraulic fracturing<br />

• Well integrity pressure testing<br />

• Fluid volumes injected during well stimulation and type and amount of additives and<br />

proppant used<br />

• Pressures used during hydraulic fracturing<br />

• Fracture growth data including that predicted and that observed<br />

• Flowback and produced water data following hydraulic fracturing including volume,<br />

disposition, and duration<br />

The EPA is creating queries on the extracted data that are expected to determine whether drinking<br />

water resources were protected from hydraulic fracturing operations. The results of these queries<br />

may indicate the frequency and variety of construction and fracturing techniques that could lead to<br />

impacts on drinking water resources. The results may provide, but may not be limited to,<br />

information on the following:<br />

• Sources of water used for hydraulic fracturing<br />

• Vertical distance between hydraulically fractured zones and the top of cement sheaths<br />

• Quality of cementing near hydraulic fracturing zones, as determined by a cement bond<br />

index<br />

• Number of well casing intervals left uncemented and whether there are aquifers in those<br />

intervals<br />

• Distribution of depths of hydraulically fractured zones from the surface<br />

• Frequency with which various tests are conducted, including casing shoe pressure tests<br />

and casing pressure tests<br />

• Types of rock formations hydraulically fractured<br />

• Types of well completions (e.g., vertical, horizontal)<br />

• Types and amounts of proppants and chemicals used during hydraulic fracturing<br />

• Amounts of fracture growth<br />

53


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

• Distances between wells hydraulically fractured and geologic faults<br />

• Proportions of fluid flowed back to the surface following hydraulic fracturing and the<br />

disposition of the flowback<br />

3.4.5. Next Steps<br />

Additional Database Analysis. The EPA plans to conduct further reviews of the well files to extract<br />

information relating to water acquisition for hydraulic fracturing, hydraulic fracturing fluid<br />

injection, and wastewater management.<br />

Statistical Analysis. Once the data analysis has been completed, where possible, extrapolation of the<br />

results will be performed to the sampled universe of 24,925 wells, using methods consistent with<br />

published statistical practices (Kish, 1965).<br />

Confidential Business Information. The EPA is working with the oil and gas operators to determine<br />

how best to summarize the results so that CBI is protected while upholding the agency’s<br />

commitment to transparency.<br />

3.4.6. Quality Assurance Summary<br />

The EPA and its contractor, The Cadmus Group, Inc., are evaluating the well file contents. The QAPP<br />

associated with this project, “National Hydraulic Fracturing Study Evaluation of Existing Production<br />

Well File Contents (Version 1),” was approved on January 4, 2012 (US EPA, 2012j). A supplemental<br />

QAPP developed by Cadmus was approved on March 6, 2012 (Cadmus Group Inc., 2012b). Each<br />

team involved in the well file review underwent a separate TSA by the designated EPA QA Manager<br />

to ensure compliance with the approved QAPP. The audits occurred between April and August of<br />

2012. No corrective actions were identified.<br />

Westat, under contract with the EPA, is providing statistical support for the well file analysis. A<br />

QAPP, “Quality Assurance Project Plan v1.1 for Hydraulic Fracturing,” was developed by Westat and<br />

approved on July 15, 2011 (Westat, 2011).<br />

3.5. FracFocus Analysis<br />

3.5.1. Relationship to the Study<br />

Extracting data from FracFocus allows the EPA to gather publicly available, nationwide information<br />

on the water volumes and chemicals used in hydraulic fracturing operations, as reported by oil and<br />

gas operating companies. Data compiled from FracFocus are being used to help inform answers to<br />

the research questions listed in Table 24.<br />

Table 24. Secondary research questions addressed by extracting data from FracFocus, a nationwide hydraulic<br />

fracturing chemical registry.<br />

Water Cycle Stage<br />

Water acquisition<br />

Chemical mixing<br />

Applicable Research Questions<br />

How much water is used in hydraulic fracturing operations, and what<br />

are the sources of this water<br />

What are the identities and quantities of chemicals used in hydraulic<br />

fracturing fluids, and how might this composition vary at a given site<br />

and across the country<br />

54


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

3.5.2. Project Introduction<br />

At the time the draft study plan was written in early 2011, the Ground Water Protection Council<br />

and the Interstate Oil and Gas Compact Commission jointly launched a new national registry for<br />

chemicals used in hydraulic fracturing, called FracFocus (http://www.fracfocus.org; (GWPC,<br />

2012b)). This registry, which has become widely accepted as the national hydraulic fracturing<br />

chemical registry, is an online repository where oil and gas well operators can upload information<br />

regarding the chemical compositions of hydraulic fracturing fluids used in specific oil and gas<br />

production wells. It has become one of the largest sources of data and information on chemicals<br />

used in hydraulic fracturing and may be the largest single source of publicly disclosed data for these<br />

chemicals. The registry also contains information on well locations, well depth, and water use.<br />

Confidential business information is not disclosed in FracFocus to protect proprietary or sensitive<br />

information.<br />

FracFocus began as a voluntary program on January 1, 2011. Since its introduction, the amount of<br />

data in FracFocus has been steadily increasing. As of May 2012, the registry contained information<br />

on nearly 19,000 wells for which hydraulic fracturing fluid disclosures were entered (GWPC,<br />

2012b). Seven states require operators to use FracFocus to report the chemicals used in hydraulic<br />

fracturing operations. In addition, many states are expected to pass or are working on legislation to<br />

require reporting with FracFocus. 37<br />

Although it represents neither a random sample nor a complete representation of the wells<br />

fractured during this time period, the number of well disclosures in FracFocus may constitute a<br />

large portion of the number of wells hydraulically fractured in the United States for this time<br />

period. For comparison, nine hydraulic fracturing service companies reported that nearly 25,000<br />

wells were fractured between September 2009 and October 2010, as described in Section 3.3.<br />

This analysis is gathering information on water and chemical use in hydraulic fracturing operations<br />

and attempts to answer the following questions:<br />

• What are the patterns of water usage in hydraulic fracturing operations reported in<br />

FracFocus<br />

• What are the different sources of water reported in FracFocus, and is it possible to<br />

determine the relative proportions by volume or mass of these different sources of water<br />

• What are the identities of chemicals used in hydraulic fracturing fluids reported in<br />

FracFocus<br />

• Which chemicals are reported most often in FracFocus<br />

• What is the geographic distribution of the most frequently reported chemicals in<br />

FracFocus<br />

37 The seven states requiring disclosure to FracFocus are Colorado, Louisiana, Montana, North Dakota, Oklahoma,<br />

Pennsylvania, and Texas. As of September 2012, the EPA is aware of eight more states considering the use of FracFocus:<br />

Alaska, California, Illinois, Kansas, Kentucky, New Mexico, Ohio, and West Virginia.<br />

55


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

3.5.3. FracFocus Data<br />

All data in FracFocus are entered by oil and gas companies that have agreed to “disclose the<br />

information in the public interest” (GWPC, 2012b). The Ground Water Protection Council, the<br />

organization that administers the registry, makes no specific claim about data quality in FracFocus.<br />

There is considerable variability in the posted data because they are uploaded by many different<br />

companies, including operator and service companies. Although FracFocus uses some built-in QA<br />

checks during the data upload process, several data quality issues are not addressed by these<br />

protocols. As a result, the EPA conducted a QA review of the data, as described in the next section.<br />

Data in FracFocus are presented in individual PDF formats for individual wells; an example PDF is<br />

provided in Figure 13. Individual wells can be searched using a Google Maps application<br />

programming interface. In addition, well disclosure records can be searched by state, county, and<br />

operator. Results are returned by listing links to individual PDF files. Because only single well<br />

disclosure records are downloadable, systematic analysis of larger datasets is more challenging.<br />

Data must be extracted and transformed into more appropriate formats (e.g., a Microsoft Access<br />

database) for this type of analysis.<br />

Data in FracFocus can be classified into two general types: well or “header” data and chemical- or<br />

ingredient-specific data. Header data describe information about each well, including the fracture<br />

date, API number, operator, well location, and total fluid volume, as shown in Figure 13. Chemicalspecific<br />

data provide the trade names of ingredients, the chemicals found in these ingredients, and<br />

the concentrations used in the hydraulic fracturing fluid. Some well disclosures include information<br />

on the type or source of water in the chemical-specific data table.<br />

The EPA has downloaded data in FracFocus on wells hydraulically fractured during 2011 and the<br />

beginning of 2012. It is beyond the scope of this project to evaluate the quality or<br />

representativeness on a national scale of the data submitted to FracFocus by oil and gas operators.<br />

The data cannot be assumed to be a complete or statistically representative of all hydraulically<br />

fractured wells. However, because FracFocus contains several thousands of well disclosures<br />

distributed throughout the United States, the EPA believes that the data in FracFocus are generally<br />

indicative of hydraulic fracturing activities during the time period covered. Therefore, it may be<br />

possible to find geographic patterns of occurrence or usage, including volume of water, frequency<br />

of chemical usage, and amounts of chemicals used, assuming that data in FracFocus meet quality<br />

requirements.<br />

56


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Figure 13. Example of data disclosed through FracFocus. Data included in each PDF can be classified into two general types: well or “header” data and chemical- or<br />

ingredient-specific data. Header data are located in the top table, and ingredient-specific data are found in the bottom table. Provided by Ground Water Protection<br />

Council.<br />

57


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

3.5.4. Research Approach<br />

Data were first extracted from the FracFocus website, put into more appropriate formats for QA<br />

review, and then organized into a final database for analysis of fracturing fluid chemicals and water<br />

usage and source. The geographic coordinates provided for wells will be linked to both the chemical<br />

and water data (Figure 13) to determine if regional patterns exist. A QA review was performed<br />

following the data extraction and initial processing. The last stage of this project involves the<br />

quantitative analyses of the QA-reviewed data. These three stages are described in more detail<br />

below.<br />

3.5.4.1. Data Extraction and Organization<br />

Records for 12,306 wells hydraulically fractured from January 1, 2011, through February 27, 2012,<br />

were extracted from FracFocus PDF files and converted to XML using Adobe Acrobat Pro X<br />

software. Header- and chemical-specific data were mined from the XML files using text recognition<br />

software (Cadmus Group Inc., 2012b). 38 Using this technique, data representing 12,173 (>98% of<br />

the downloaded records) well records were compiled. Once fully processed, the data records were<br />

organized into two working files: one file containing header data that included well-specific<br />

geography, fracturing fluid volume, and well depth and one file containing chemical-specific data.<br />

The working files are linked by unique well identification numbers assigned by the contractor that<br />

developed the database for EPA.<br />

3.5.4.2. Data Quality Assurance Review<br />

Manual and automated methods were used to assess the data quality and make necessary<br />

adjustments. Records in the header data working file were flagged according to the following<br />

criteria: duplicate records, as identified by identical API numbers; fracture dates outside the<br />

January 1, 2011, to February 27, 2012, time period; anomalously large or small volumes of water;<br />

and anomalously deep or shallow true vertical depths. These records were kept in the working files,<br />

but flagged in order to exclude them from future analyses. Half of the duplicate records were<br />

excluded from all queries and analyses.<br />

Spatial data from the well records include three sources, which can be used to perform quality<br />

checks: state and county names, latitude and longitude coordinates, and the state and county<br />

information encoded in the first five digits of the API Well Number (Figure 13). To validate the<br />

location of the wells, the state and county information from each of the locational fields was<br />

compared. State and county information (ESRI, 2010a, b) was assigned to the latitude and<br />

longitude coordinates by spatially joining the data in ArcGIS (ESRI, version 10). Validated spatial<br />

location was available for 12,163 wells (>99% of records extracted) (Cadmus Group Inc., 2012b).<br />

Chemical names in the “Ingredients” field of chemical-specific data table were standardized<br />

according to the CASRN provided in the associated “Chemical Abstract Service Number” field<br />

38 The text recognition software is highly sensitive to inconsistencies in reporting. If an operator departs from the general<br />

template when creating the well record, the record will be passed over or data will be extracted incorrectly. The<br />

contractor was able to convert data from 12,173 of the 12,302 well records into a more useable format (Cadmus Group<br />

Inc., 2012b).<br />

58


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

(Figure 13). As described in Chapter 6, the EPA has compiled and curated a list of chemicals<br />

reported to be used in hydraulic fracturing fluids from many data sources. This list was used to<br />

standardize the chemical names provided in FracFocus by matching CASRNs. 39<br />

Water sources were also identified from the “Ingredients” field. Data were first organized to<br />

identify wells where water has been listed as a trade name or ingredient and has been used as a<br />

“carrier” or “base” fluid, excluding records that indicated the water has been used as a solvent for<br />

hydraulic fracturing chemicals. Additionally, records listing the CASRN for water (7732-18-5) and<br />

an additive concentration of 70% to 100% were identified.<br />

3.5.4.3. Data Analysis<br />

Following the QA review, all data were organized into four data tables: locational data for each well<br />

disclosure, the original chemical-specific data for each well disclosure, the QA-reviewed chemicalspecific<br />

data for each well disclosure, and records with water as ingredient. These four tables have<br />

been imported into a database and linked together using key fields, where they can be used for the<br />

analyses described below. The raw, pre-QA data values for well disclosures and chemical<br />

ingredients as they were exported from FracFocus have also been imported into the database for<br />

baseline reference data to prevent any loss of original operator data.<br />

Water Acquisition. Total water volume data that meet the QA requirements are being used to<br />

analyze general water usage patterns on national, state, and county scales of interest. Additional<br />

queries may be run that analyze water usage by operator and by production type (oil or gas).<br />

Data will be summarized by water source or type for records where this information is provided.<br />

Concentrations of water by source type are generally found in the “Maximum Ingredient<br />

Concentration in HF (hydraulic fracturing) Fluid” field (Figure 13), which is reported as a<br />

percentage by mass, not percentage by total water volume. In some situations, there will be enough<br />

i<br />

information in FracFocus to calculate water volumes by type (V H2O ), whether fresh water (e.g.,<br />

surface water) or non-fresh water (e.g. recycled/produced, saline, seawater or brine). Given the<br />

FracFocus-reported total water volume (V total H2O ) (US gallons) and assuming that volumes are<br />

effectively additive, and where n is the number of water types,<br />

total<br />

V ≅ ∑ n i<br />

H2O i=1 V H2O (1)<br />

using the FracFocus-reported maximum water concentration in the hydraulic fracturing fluid<br />

i<br />

(percent by mass for each water type, x H2O ), and assuming an average density for each water type<br />

i<br />

(ρ H2O ) (lb/US gallons), the volume of each water type is expressed as:<br />

i<br />

i<br />

x H2O<br />

ρ H2O<br />

VH2O = i m total (i = 1, n) (2)<br />

With n equations and n unknowns represented by equations (1) and (2), the unknown total mass of<br />

the hydraulic fracturing fluid (m total) (lb) can be calculated:<br />

39 CASRNs not already found on the EPA’s list of chemicals reported to be used in hydraulic fracturing fluids were added<br />

to the list following the process outlined in Chapter 6.<br />

59


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

total<br />

V H2O<br />

mtotal = i (3)<br />

∑n<br />

x H2O<br />

i=1<br />

ρ i<br />

H2O<br />

i<br />

and the volume of each water type (V H2O ) back-calculated using equation (2). 40<br />

This calculation can only be made in the situation where the density of the fluid is known or<br />

reported. For example, in the situation where a FracFocus ingredient is clearly labeled fresh<br />

(surface) water and carrier or base fluid, a water density may be assumed between 8.34 lb/US<br />

gallon at 32 °F and 8.24 lb/US gallon at 100 °F (Lide, 2008). In other situations, the density for the<br />

carrier or base fluid may be reported in the FracFocus comment field.<br />

Chemical Usage. Queries of the FracFocus data will include the total number of unique chemical<br />

records nationally, by state, per production type (oil or gas), fracture date, and operator<br />

represented. Additionally, the data may be queried to identify the frequency or number of well<br />

disclosures in which each chemical is used nationally, by state, per production type, within a<br />

fracture date range, and by operator represented. Lists of the top 20 to 30 most frequently used<br />

chemicals in hydraulic fracturing are likely to be generated at the nation, region, or state level.<br />

Some of the most frequently occurring chemicals will be mapped to show distribution of<br />

occurrence. Since chemicals claimed as CBI or proprietary do not have to be reported in FracFocus,<br />

the number of chemicals disclosed is likely to be lower than the total number of chemicals used.<br />

3.5.5. Status and Preliminary Data<br />

The data have been extracted from FracFocus, reviewed for quality issues, and organized in a<br />

database for analysis. Draft queries have been developed for water usage and chemical frequency<br />

occurrence nationwide using the database. Preliminary analyses have been conducted as of<br />

November 2012. Table 25 summarizes, by state, the well data that were downloaded from<br />

FracFocus in early 2012.<br />

Table 25. Number of wells, by state, with data in FracFocus as of February 2012. These data represent wells<br />

fractured and entered into FracFocus between January 1, 2011, and February 27, 2012.<br />

State<br />

Number of Wells<br />

Alabama 54<br />

Alaska 24<br />

Arkansas 807<br />

California 79<br />

Colorado 2,307<br />

Kansas 22<br />

Louisiana 621<br />

Mississippi 1<br />

Montana 28<br />

New Mexico 421<br />

State<br />

Number of Wells<br />

North Dakota 359<br />

Ohio 11<br />

Oklahoma 414<br />

Pennsylvania 1,050<br />

Texas 4,859<br />

Utah 409<br />

Virginia 23<br />

West Virginia 93<br />

Wyoming 591<br />

Total 12,173<br />

40 The EPA recognizes that volume is not a conserved quantity and estimates that the error introduced by assuming that<br />

volumes are additive is, in this case, negligible when compared to expected volume and density reporting errors.<br />

60


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

During the QA review of the data, the EPA identified 422 pairs of potential duplicate well disclosure<br />

records (844 total records). A total of 277,029 chemicals were reported in all of the well disclosure<br />

records. This number includes chemicals listed multiple times (either for the same well or in many<br />

wells) and 12,464 instances where “water” was listed as an ingredient in the chemical-specific data<br />

table. The QA review of the chemicals identified 347 unique ingredients that match the EPA CASRN<br />

list of chemicals and approximately 60 CASRNs that were not previously known to be used in<br />

hydraulic fracturing fluids. One hundred eighty-four well records had ingredient lists that fully<br />

matched the EPA CASRN list. Chemical entries in FracFocus that contained “CBI,” “proprietary,” or<br />

“trade secret” as an ingredient were only 1.3% (3,534 of 277,029) of all chemical ingredients<br />

reported in FracFocus. Operators reported at least one chemical ingredient as “CBI,” “proprietary,”<br />

or “trade secret” in 1,924 well records.<br />

Water was identified as a carrier or base fluid in 10,700 well records (88% of the 12,173 well<br />

records successfully extracted from FracFocus). Seven categories of source water were identified:<br />

fresh, surface, sea, produced, recycled, brine, and treated. Definitions for the categories are not<br />

provided by operators or FracFocus and some categories appear to overlap or may be synonymous.<br />

Only 1,484 well records identified a water source for those wells that used water as a carrier or<br />

base fluid.<br />

3.5.6. Next Steps<br />

The EPA will complete its analysis of the FracFocus data that have already been downloaded. In<br />

addition, the EPA plans to complete another data download in order to obtain a second year’s<br />

worth of data. Once the second round of data has been extracted, the EPA will conduct a QA review<br />

and data analysis similar to the one described for the first round of downloaded data.<br />

3.5.7. Quality Assurance Summary<br />

The EPA and its contractor, The Cadmus Group, Inc., are extracting and analyzing data from<br />

FracFocus. The QAPP associated with this project, “Analysis of Data Extracted from FracFocus<br />

(Version 1),” was approved in early August 2012 (US EPA, 2012g). A TSA of the analysis was<br />

conducted by the designated EPA QA Manager shortly after on August 15, 2012; no corrective<br />

actions were identified. A supplemental QAPP developed by Cadmus was approved March 6, 2012<br />

(Cadmus Group Inc., 2012b).<br />

61


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

4. Scenario Evaluations<br />

The objective of this approach is to use computer models to explore hypothetical scenarios across<br />

the hydraulic fracturing water cycle. The models include models of generic engineering and<br />

geological scenarios and, where sufficient data are available, models of site-specific or regionspecific<br />

characteristics. This chapter includes progress reports for the following projects:<br />

4.1. Subsurface Migration Modeling....................................................................................................................... 62<br />

Numerical modeling of subsurface fluid migration scenarios that explore the potential for<br />

gases and fluids to move from the fractured zone to drinking water aquifers<br />

4.2. Surface Water Modeling...................................................................................................................................... 75<br />

Modeling of concentrations of selected chemicals at public water supplies downstream from<br />

wastewater treatment facilities that discharge treated hydraulic fracturing wastewater to<br />

surface waters<br />

4.3. Water Availability Modeling.............................................................................................................................. 80<br />

Assessment and modeling of current and future scenarios exploring the impact of water usage<br />

for hydraulic fracturing on drinking water availability in the Upper Colorado River Basin and<br />

the Susquehanna River Basin<br />

4.1. Subsurface Migration Modeling<br />

Lawrence Berkeley National Laboratory (LBNL), in consultation with the EPA, will simulate the<br />

hypothetical subsurface migration of fluids (including gases) resulting from six possible<br />

mechanisms using computer models. The selected mechanisms address the research questions<br />

identified in Table 26.<br />

Table 26. Secondary research questions addressed by simulating the subsurface migration of gases and fluids<br />

resulting from six possible mechanisms.<br />

Water Cycle Stage<br />

Well injection<br />

Applicable Research Questions<br />

• How effective are current well construction practices at<br />

containing gases and fluids before, during, and after fracturing<br />

• Can subsurface migration of fluids or gases to drinking water<br />

resources occur and what local geologic or man-made features<br />

may allow this<br />

4.1.1. Project Introduction<br />

Stakeholders have expressed concerns about hydraulic fracturing endangering subsurface drinking<br />

water resources by creating high permeability transport pathways that allow hydrocarbons and<br />

other fluids to escape from hydrocarbon-bearing formations (US EPA, 2010b, d, e, f, g). Experts<br />

continue to debate the extent to which subsurface pathways could cause significant adverse<br />

consequences for ground water resources (Davies, 2011; Engelder, 2012; Harrison, 1983, 1985;<br />

Jackson et al., 2011; Myers, 2012a, b; Osborn et al., 2011; Warner et al., 2012). The segment of the<br />

population that receives drinking water from private wells may be especially vulnerable to health<br />

impacts from impaired drinking water. Unlike water distributed by public water systems, water<br />

62


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

from private drinking water wells is not subject to National Primary Drinking Water Regulations,<br />

and water quality testing is at the discretion of the well owner.<br />

Lawrence Berkeley National Laboratory, in coordination with the EPA, is using numerical<br />

simulations to investigate six possible mechanisms that could lead to upward migration of fluids,<br />

including gases, from a shale gas reservoir and the conditions under which such hypothetical<br />

scenarios may be possible. The possible mechanisms include:<br />

• Scenario A (Figure 14): Defective or insufficient well construction coupled with excessive<br />

pressure during hydraulic fracturing operations results in damage to well integrity during<br />

the stimulation process. A migration pathway is then established through which fluids<br />

could travel through the cement or area near the wellbore into overlying aquifers. In this<br />

scenario, the overburden is not necessarily fractured.<br />

• Scenario B1 (Figure 15): Fracturing of the overburden because inadequate design of the<br />

hydraulic fracturing operation results in fractures allowing fluid communication, either<br />

directly or indirectly, between shale gas reservoirs and aquifers above them. Indirect<br />

communication would occur if fractures intercept a permeable formation between the<br />

shale gas formation and the aquifer. Generally, the aquifer would be located at a more<br />

shallow depth than the permeable formation.<br />

• Scenario B2 (Figure 16): Similar to Scenario B1, fracturing of the overburden allows<br />

indirect fluid communication between the shale gas reservoir and the aquifers after<br />

intercepting conventional hydrocarbon reservoirs, which may create a dual source of<br />

contamination for the aquifer.<br />

• Scenario C (Figure 17): Sealed/dormant fractures and faults are activated by the hydraulic<br />

fracturing operation, creating pathways for upward migration of hydrocarbons and other<br />

contaminants.<br />

• Scenario D1 (Figure 18): Fracturing of the overburden creates pathways for movement of<br />

hydrocarbons and other contaminants into offset wells (or their vicinity) in conventional<br />

reservoirs with deteriorating cement. The offset wells may intersect and communicate<br />

with aquifers, and inadequate or failing completions/cement can create pathways for<br />

contaminants to reach the ground water aquifer.<br />

• Scenario D2 (Figure 19): Similar to Scenario D1, fracturing of the overburden results in<br />

movement of hydrocarbons and other contaminants into improperly closed offset wells<br />

(or their vicinity) with compromised casing in conventional reservoirs. The offset well<br />

could provide a low-resistance pathway connecting the shale gas reservoir with the<br />

ground water aquifer.<br />

The research focuses on hypothetical causes of failure related to fluid pressure/flow and<br />

geomechanics (as related to operational and geological conditions and properties), and does not<br />

extend to investigations of strength of casing and tubing materials (an area that falls within the<br />

confines of mechanical engineering). Damage to the well casing due to corrosive reservoir fluids<br />

was one other scenario originally considered. Corrosion modeling requires a detailed chemical<br />

engineering analysis that is beyond the scope of this project, which focuses on geophysical and<br />

63


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

mechanical scenarios, so it is not a scenario pursued for this project. Additionally, hypothetical<br />

scenarios that would cause failure of well structural integrity (e.g., joint splits) are an issue beyond<br />

the scope of this study, as they involve material quality and integrity, issues not unique to hydraulic<br />

fracturing.<br />

Figure 14. Scenario A of the subsurface migration modeling project. This scenario simulates a hypothetical migration<br />

pathway that occurs when a defective or insufficiently constructed well is damaged during excessive pressure from<br />

hydraulic fracturing operations. A migration pathway is established through which fluids could travel through the<br />

cement or area near the wellbore into overlying ground water aquifers.<br />

64


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Figure 15. Scenario B1 of the subsurface migration modeling project. This hypothetical scenario simulates fluid<br />

communication, either directly or indirectly, between shale gas reservoirs and ground water aquifers as a result of the<br />

hydraulic fracturing design creating fractures in the overburden.<br />

65


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Figure 16. Scenario B2 of the subsurface migration modeling project. Similar to B1, this hypothetical scenario<br />

simulates fluid communication, either directly or indirectly, between shale gas reservoirs and ground water aquifers<br />

as a result of the hydraulic fracturing design creating fractures in the overburden. The fractures intercept a<br />

conventional oil/gas reservoir before communicating with the ground water aquifer, which may create a dual source of<br />

contamination in the aquifer.<br />

66


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Figure 17. Scenario C of the subsurface migration modeling project. This hypothetical scenario simulates upward<br />

migration of hydrocarbons and other contaminants through sealed/dormant fractures and faults activated by the<br />

hydraulic fracturing operation.<br />

67


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Figure 18. Scenario D1 of the subsurface migration modeling project. This hypothetical scenario simulates<br />

movement of hydrocarbons and other contaminants into offset wells in conventional oil/gas reservoirs with<br />

deteriorating cement due to fracturing of the overburden. The offset wells may intersect and communicate with<br />

aquifers, and inadequate or failing completions/cement can create pathways for contaminants to reach ground water<br />

aquifers.<br />

68


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Figure 19. Scenario D2 of the subsurface migration modeling project. Similar to Scenario D1, this hypothetical<br />

scenario simulates movement of hydrocarbons and other contaminants into offset wells in conventional oil/gas<br />

reservoirs due to fracturing of the overburden. The offset wells in Scenario D2 are improperly closed with<br />

compromised casing, which provides a low-resistance pathway connecting the shale gas reservoir with the ground<br />

water aquifer.<br />

69


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

4.1.2. Research Approach<br />

Objectives of the subsurface migration scenario evaluation research project include:<br />

• Determining whether the hypothetical migration mechanisms shown in Figures 14<br />

through 19 are physically and geomechanically possible during field operations of<br />

hydraulic fracturing and, if so, identifying the range of conditions under which fluid<br />

migration is possible.<br />

• Exploring how contaminant type, fluid pressure, and local geologic properties control<br />

hypothetical migration mechanisms and affect the possible emergence of contaminants in<br />

an aquifer.<br />

• Conducting a thorough analysis of sensitivity to the various factors affecting contaminant<br />

transport.<br />

• Assessing the potential impacts on drinking water resources in cases of fluid migration.<br />

This research project does not assess the likelihood of a hypothetical scenario occurring during<br />

actual field operations.<br />

Computational Codes. The LBNL selected computational codes able to simulate the flow and<br />

transport of gas, water, and dissolved contaminants concurrently in fractures and porous rock<br />

matrices. The numerical models used in this research project couple flow, transport,<br />

thermodynamics, and geomechanics to produce simulations to promote understanding of<br />

conditions in which fluid migration occurs.<br />

Simulations of contaminant flow and migration began in December 2011 and identified a number of<br />

important issues that significantly affected the project approach. More specifically, the numerical<br />

simulator needed to include the following processes in order to accurately describe the<br />

hypothetical scenario conditions:<br />

• Darcy and non-Darcy (Forchheimer or Barree and Conway) flow through the matrix and<br />

fractures of fractured media<br />

• Inertial and turbulent effects (Klinkenberg effects)<br />

• Real gas behavior<br />

• Multi-phase flow (gas, aqueous, and potentially an organic phase of immiscible substances<br />

involved in the hydraulic fracturing process)<br />

• Density-driven flow<br />

• Mechanical dispersion, in addition to advection and molecular diffusion<br />

• Sorption (primary and secondary) of ions introduced in hydraulic fracturing-related<br />

processes and gases onto the grains of the porous media, involving one of three possible<br />

sorption models (linear, Langmuir, or Freundlich) under equilibrium or kinetic conditions<br />

70


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Thermal differentials between ground water and shale gas reservoirs are substantial and may<br />

significantly impact contaminant transport processes. Thus, the simulator needed to be able to<br />

account for the following processes in order to fully describe the physics of the problem:<br />

• Coupled flow and thermal effects, which affect fluid viscosity, density, and buoyancy and,<br />

consequently, the rate of migration.<br />

• Effect of temperature on solubility. Lower temperatures can lead to supersaturation of<br />

dissolved gases or dissolved solids. The latter can result in halite formation stemming<br />

from salt precipitation, caused by lower temperatures and pressures as naturally<br />

occurring brines ascend toward the ground water. Halite precipitation can have a<br />

pronounced effect on both the specific fractures and the overall matrix permeability.<br />

There is currently no single numerical model that includes all of these processes. Thus, the LBNL<br />

chose the Transport of Unsaturated Groundwater and Heat (TOUGH) family of codes 41 (Moridis et<br />

al., 2008) in combination with the existing modules listed in Table 27 to create a model that better<br />

simulates the subsurface flow and geomechanical conditions encountered in the migration<br />

scenarios.<br />

Table 27. Modules combined with the Transport of Unsaturated Groundwater and Heat (TOUGH) (Moridis et al.,<br />

2008) family of codes to create simulations of subsurface flow and geomechanical conditions encountered in the<br />

migration scenarios designed by Lawrence Berkeley National Laboratory.<br />

Module<br />

Purpose<br />

Describes the coupled flow of a real gas mixture and heat in geologic<br />

TOUGH+Rgas*<br />

media<br />

Describes the non-isothermal two-phase flow of a real gas mixture and<br />

TOUGH+RgasH2O* water and the transport of heat in a gas reservoir, including tight/shale<br />

gas reservoirs<br />

TOUGH+RGasH2OCont † Describes physics and chemistry of flow and transport of heat, water,<br />

gases, and dissolved contaminants in porous/fractured media<br />

Simulates geomechanical behavior of multiple porosity/permeability<br />

ROCMECH §<br />

continuum systems and can accurately simulate the evolution and<br />

propagation of fractures in a formation following hydraulic fracturing<br />

* (Moridis and Freeman, 2012)<br />

†<br />

(Moridis and Webb, 2012)<br />

§<br />

(Kim and Moridis, 2012a, b, c, d, e)<br />

The TOUGH+ code includes equation-of-state modules that describe the non-isothermal flow of real<br />

gas mixtures, water, and solutes through fractured porous media and accounts for all processes<br />

involved in flow through tight and shale gas reservoirs (i.e., gas-specific Knudsen diffusion, gas and<br />

solute sorption onto the media, non-Darcy flow, salt precipitation as temperature and pressure<br />

drop in the ascending reservoir, etc.) (Freeman, 2010; Freeman et al., 2011; Freeman et al., 2009a,<br />

b; Freeman et al., 2012; Moridis et al., 2010; Olorode, 2011). The LBNL paired relevant modules<br />

with TOUGH+ code: one code, TOUGH+RGasH2OCont (Moridis and Freeman, 2012), addresses the<br />

41 The TOUGH codes include TOUGH2, T2VOC, TMVOC, TOUGH2-MP, TOUGHREACT, TOUGH+, AND iTOUGH2. More<br />

information on the codes can be found at http://esd.lbl.gov/research/projects/tough.<br />

71


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

physics and chemistry of flow and transport of heat, water, gases, and dissolved contaminants in<br />

porous/fractured media; a second code, TOUGH+RgasH2O (Moridis and Webb, 2012), describes the<br />

coupled flow of a gas mixture and water and the transport of heat; a third code, TOUGH+Rgas<br />

(Moridis and Webb, 2012), is limited to the coupled flow of a real gas mixture and heat in geologic<br />

media.<br />

A geomechanical model, ROCMECH, was also coupled with the TOUGH+ code and modules (Table<br />

27) and describes the interdependence of flow and geomechanics including fracture growth and<br />

propagation (Kim and Moridis, 2012a, b, c, d, e). The ROCMECH 42 code is designed for the rigorous<br />

analysis of either pure geomechanical problems or, when fully coupled with the TOUGH+ multiphase,<br />

multi-component, non-isothermal code, for the simulation of the coupled flow and<br />

geomechanical system behavior in porous and fractured media, including activation of faults and<br />

fractures. The coupled TOUGH+ ROCMECH codes allow the investigation of fracture growth during<br />

fluid injection of water (after their initial development during hydraulic fracturing) using fully<br />

dynamically coupled flow and geomechanics and were used in a series of fracture propagation<br />

studies (Kim and Moridis, 2012a, b, c, d, e). The ROCMECH code developed by the LBNL for this<br />

study includes capabilities to describe both tensile and shear failure based on the Mohr-Coulomb<br />

model, multiple porosity concepts, non-isothermal behavior, and transverse leak-off (Kim and<br />

Moridis, 2012a).<br />

Input Data. Input data supporting the simulations are being estimated using information from the<br />

technical literature, data supplied by the EPA, and expert judgment. Input data include:<br />

• Site stratigraphy<br />

• Rock properties (grain density, intrinsic matrix permeability, permeability of natural<br />

fracture network, matrix and fracture porosity, fracture spacing and aperture)<br />

• Initial formation conditions (fracture and matrix saturation, pressures)<br />

• Gas composition<br />

• Pore water composition<br />

• Gas adsorption isotherm<br />

• Thermal conductivity and specific heat of rocks<br />

• Parameters for relative permeability<br />

• Hydraulic fracturing pressure<br />

• Number of hydraulic fracturing stages<br />

• Injected volumes<br />

42 ROCMECH is based on an earlier simulator called ROCMAS (Noorishad and Tsang, 1997; Rutqvist et al., 2001). The<br />

ROCMECH simulator employs the finite element method, includes several plastic models such as the Mohr-Coulomb and<br />

Drucker-Prager models, and can simulate the geomechanical behavior of multiple porosity/permeability continuum<br />

systems. Furthermore, ROCMECH can accurately simulate the process of hydraulic fracturing, i.e., the evolution and<br />

propagation of fractures in the formation following stimulation operations.<br />

72


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

• Pressure evolution during injection<br />

• Volumes of fracturing fluid recovered<br />

Uncertainty in the data will be addressed by first analyzing base cases that involve reasonable<br />

estimates of the various parameters and conditions and then conducting sensitivity analyses that<br />

cover (and extend beyond) the possible range of expected values of all relevant parameters.<br />

4.1.3. Status and Preliminary Data<br />

The subsurface migration modeling project is proceeding along two main tracks. The first<br />

addresses the geomechanical reality of the mechanisms and seeks to determine whether it is<br />

physically possible (as determined and constrained by the laws of physics and the operational<br />

quantities and limitations involved in hydraulic fracturing operations) for the six migration<br />

mechanisms (Scenarios A to D2) to occur. The second axis focuses on contaminant transport,<br />

assuming that a subsurface migration has occurred as described in the six scenarios, and attempts<br />

to determine a timeframe for contaminants (liquid or gas phase) escaping from a shale gas<br />

reservoir to reach the ground water aquifer.<br />

Analysis of Consequences of Geomechanical Wellbore Failure (Scenario A). A large database of<br />

relevant publications has been assembled, and several important well design parameters and<br />

hydraulic fracturing operational conditions have been identified as a foundation for the simulation.<br />

Two pathways for migration have been considered using TOUGH+RGasH2OCont: cement<br />

separation from the outer casing or a fracture pattern affecting the entire cement, from the<br />

producing formation to the point where the well intercepts the ground water formation.<br />

A separate geomechanical study using TOUGH+RealGasH2O and ROCMECH will also assess the<br />

feasibility of either a fracture developing in weak cement around a wellbore or a cement-wellbore<br />

separation during the hydraulic fracturing process. The numerical simulation of the fracture<br />

propagation considered fracture development in the cement near the “heel” of a horizontal well<br />

during stimulation immediately after creation of the first fracture using varied geomechanical<br />

properties of gas-bearing shales. The work also involves sensitivity analyses of factors that are<br />

known to be important, as well as those that appear to have secondary effects (for completeness).<br />

Recent activities have focused mainly on such sensitivity analyses.<br />

Analysis of the Consequences of Induced Fractures Reaching Ground Water Resources and after<br />

Intercepting Conventional Reservoirs (Scenarios B1 and B2). A high-definition geomechanical study,<br />

involving a complex fracture propagation model that incorporates realistic data and parameters (as<br />

gleaned from the literature and discussions with industry practitioners) was completed. A<br />

sensitivity analysis of the fracture propagation to the most important geomechanical properties<br />

and conditions is partially completed and will be included in the final publication.<br />

73


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Simulations of gas and contaminant migration from the shale gas reservoir through fractures into<br />

ground water are also in progress. The simulation domain is subdivided up to 300,000 elements 43<br />

and up to 1.2 million equations, which requires very long execution times that can range from<br />

several days to weeks. Work continues to streamline the processing of the simulation to<br />

significantly reduce the execution time requirements.<br />

Scoping calculations are in development to provide time estimates for the migration of gas and<br />

dissolved contaminants from the shale gas reservoir to the drinking water resource through a<br />

connecting fracture. As illustrated in Figure 15, the simulated system is composed of a 100-meterthick<br />

aquifer (from 100 to 200 meters below the surface), a fracture extending from the bottom of<br />

the gas reservoir at 1,200 meters below surface to the base of the aquifer, which is 1,000 meters<br />

above the gas reservoir. These scoping studies indicated that the most important parameters and<br />

conditions were the permeability of the gas reservoir (matrix), the fracture permeability, the<br />

distance between the aquifer and the shale reservoir, and the pressure regimes in the aquifer and<br />

the shale. Results from this work are being analyzed and will be published when complete.<br />

Analysis of Consequences of Activation of Native Faults and Fractures (Scenario C). The simulation<br />

conditions for the analysis of contaminant transport through native fractures and faults in response<br />

to the stimulation process have been determined, and the variations used to conduct a sensitivity<br />

analysis are being developed.<br />

A geomechanical study using the TOUGH-FLAC 44 simulator began in March 2012 to investigate the<br />

possibility that hydraulic fracturing injections may create a pathway for transport through fault<br />

reactivation. The simulation input represents the conditions in the Marcellus Shale. Scoping<br />

calculations were developed to study the potential for injection-induced fault reactivation<br />

associated with shale gas hydraulic fracturing operations. From these scoping calculations, the<br />

LBNL simulation results suggest that the hydraulic fracturing stimulation, under conditions<br />

reported in published literature, does not appear to activate fault rupture lengths greater than 40 to<br />

50 meters and could only give rise to microseismicity (magnitude


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

through the shale gas reservoir into the weak/fractured cement around, or the unplugged wellbore<br />

of offset wells (Figures 18 and 19). The LBNL is investigating two mechanisms for fluid<br />

communication. In the first case, the fractures extend across the shale stratum into a nearby<br />

depleted conventional reservoir with abandoned defective wells in the overburden or<br />

underburden. The energy for the lift of contaminants in this case is most likely provided by the<br />

higher pressure of the fluids in the shale (as the abandoned reservoir pressure is expected to be<br />

low) and by buoyancy; the main contaminant reaching the ground water is expected to be gas. In<br />

the second case, fractures extend from a deeper over-pressurized saline aquifer through the entire<br />

thickness of the shale to an overburden (a depleted conventional petroleum reservoir with<br />

abandoned unsealed wells). The energy for the lift of contaminants in this case is most likely<br />

provided by the higher pressure of the fluids in the shale and in the saline aquifer in addition to<br />

buoyancy, and the contaminants reaching the ground water are expected to include gas and solutes<br />

encountered in the saline aquifer.<br />

4.1.4. Quality Assurance Summary<br />

The QAPP, “Analysis of Environmental Hazards Related to Hydrofracturing (Revision: 0),” was<br />

accepted by the EPA on December 7, 2011 (LBNL, 2011).<br />

A TSA of the work being performed by the LBNL was conducted on February 29, 2012. The<br />

designated EPA QA Manager found the methods in use satisfactory and further recommendations<br />

for improving the QA process were unnecessary. Work performed and scheduled to be performed<br />

was within the scope of the project. Work is proceeding on Scenarios A through D2 as described in<br />

Section 4.1.3. Reports, when presented, will be subjected to appropriate QA review.<br />

4.2. Surface Water Modeling<br />

4.2.1. Relationship to the Study<br />

The EPA is using established surface water transport theory and models to identify concentrations<br />

of selected hydraulic fracturing-relevant chemicals at public water supply intakes located<br />

downstream from wastewater treatment facilities that discharge treated hydraulic fracturing<br />

wastewater to rivers. This work is expected to provide data that will be used to answer the<br />

research question identified in Table 28.<br />

Table 28. Secondary research question addressed by modeling surface water discharges from wastewater treatment<br />

facilities accepting hydraulic fracturing wastewater.<br />

Water Cycle Stage<br />

Wastewater treatment and<br />

waste disposal<br />

Applicable Research Questions<br />

What are the potential impacts from surface water disposal of treated<br />

hydraulic fracturing wastewater on drinking water treatment<br />

facilities<br />

4.2.2. Project Introduction<br />

When an operator reduces the injection pressure applied to a well, the direction of fluid flow<br />

reverses, leading to the recovery of flowback and produced water, collectively referred to as<br />

75


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

“hydraulic fracturing wastewater.” 45 The wastewater is generally stored onsite before being<br />

transported for treatment, recycling or disposal. Most hydraulic fracturing wastewater is disposed<br />

in UIC wells. In Pennsylvania, however, wastewater has been treated in wastewater treatment<br />

facilities (WWTFs), which subsequently discharge treated wastewater to surface water bodies.<br />

The extent to which common treatment technologies used in WWTFs effectively remove chemicals<br />

found in hydraulic fracturing wastewater is currently unclear. 46 Depending in part on the<br />

concentration of chemicals in the effluent, drinking water quality and the treatment processes at<br />

public water systems (PWSs) downstream from WWTFs might be negatively affected. For example,<br />

bromide in source waters can cause elevated concentrations of brominated disinfection byproducts<br />

(DBPs) in treated drinking water (Brown et al., 2011; Plewa et al., 2008), 47 which are regulated by<br />

the National Primary Drinking Water Regulations. 48 To learn more about impacts to downstream<br />

PWSs, the Pennsylvania Department of the Environment asked 25 WWTFs that accept Marcellus<br />

wastewater to monitor effluent for parameters such as radionuclides, total dissolved solids (TDS),<br />

alkalinity, chloride, sulfate, bromide, gross alpha, radium-226 and -228, and uranium in March 2011<br />

(PADEP, 2011). The department also asked 14 PWSs with surface water intakes downstream from<br />

WWTFs that accept Marcellus wastewater to test for radionuclides, TDS, pH, alkalinity, chloride,<br />

sulfate, and bromide (PADEP, 2011). Bromide and radionuclides are of particular concern in<br />

discharges because of their carcinogenicity and reproductive and developmental affects.<br />

The EPA will use computer models—mass balance, empirical, and numerical—to estimate generic<br />

impacts of bromide and radium in wastewater discharges, based on the presence of these chemicals<br />

in discharge data from WWTFs in Pennsylvania, impacts to downstream PWSs’ ability to meet<br />

National Primary Drinking Water Regulations for DBPs and radionuclides, and the potential human<br />

health impacts from the chemicals. 49 Uranium, also a radionuclide, was frequently not detected by<br />

analytical methods for the discharges and therefore not considered for simulations. The generic<br />

model results are designed to illustrate the general conditions under which discharges might cause<br />

impacts on downstream public water supplies. The analysis will include the effect of distance to the<br />

PWS, discharge concentration, and flow rate in the stream or river, among others. The uncertainties<br />

in these quantities will be addressed through Monte Carlo analysis, as described below.<br />

A steady-state mass balance model provides an upper-bound impact assessment of the transport<br />

simulation and a partially transient approach simulates the temporal variation of effluent<br />

concentration and discharge. Key data collected to model the transport of potential contaminants<br />

include actual effluent data from WWTF discharges and receiving water body flow rates. Effluent<br />

data can be obtained from National Pollutant Discharge Elimination System (NPDES) monitoring<br />

45 Produced water is produced from many oil and gas wells and not unique to hydraulic fracturing.<br />

46 See Section 5.2 for a more thorough discussion and for EPA-funded research into this question.<br />

47 See Section 5.3 for more information on DBPs and related research. <br />

48 Authorized by the Safe Drinking Water Act.<br />

49 Discharge data for four WWTFs in Pennsylvania that accepted oil and gas wastewater during 2011 are available on the<br />

EPA’s website at http://www.epa.gov/region3/marcellus_shale/.<br />

76


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

data reported to states by the dischargers. 50 NPDES information also documents the design of the<br />

industrial treatment plants, which can give insights into the capabilities of these and similarly<br />

designed treatment plants. The US Geological Survey (USGS) provides limited water quality and<br />

flow rate data from monitoring stations within the watersheds of the receiving water bodies. The<br />

surface water modeling results will directly address the applicable secondary research question<br />

(Table 28) by evaluating the possible impacts from a permitted release of treated effluent on both a<br />

downstream drinking water intake and in a watershed where there may be multiple sources and<br />

receptors. 51<br />

4.2.3. Research Approach<br />

Multiple approaches generate results on impacts: steady-state mass balance; transient empirical<br />

modeling; and a transient, hybrid empirical-numerical model developed by the EPA. The results of<br />

the mass balance model simulate possible impacts during a large volume, high concentration<br />

discharge without natural attenuation of contaminants. The empirical model and a hybrid<br />

empirical-numerical model estimate impacts in a more realistic setting with variable chemical<br />

concentrations, discharge volumes, and flow rates of the receiving surface water. The numerical<br />

model confirms the results of the empirical and hybrid models. The numerical modeling is based on<br />

an approach developed for this study from existing methods (Hairer et al., 1991; Leonard, 2002;<br />

Schiesser, 1991; Wallis, 2007). Application of these three types of models provides a panoramic<br />

view of possible impacts and enhances confidence in the study results.<br />

Mass Balance Approach Estimates Impacts from an Upper-Bound Discharge Scenario. A simple,<br />

steady-state mass balance model simulates drinking water impacts from upper-bound discharge<br />

cases. This model assumes that the total mass of the chemical of interest is conserved during<br />

surface water transport and that the chemical concentration does not decrease due to reaction,<br />

decay, or uptake. The model estimates potential impacts to downstream PWSs using the maximum<br />

effluent concentration, maximum WWTF discharge volume, minimum flow rate in the receiving<br />

stream, and the distance to the downstream PWS intake. The EPA constructed generic discharge<br />

scenarios for rivers with varying flow regimes to determine the potential for adverse impacts at<br />

drinking water intakes. Because the parameters describing transport are uncertain, Monte Carlo<br />

techniques will be used to generate probabilistic outputs of the model.<br />

Empirical Model Estimates Impacts with Varying Discharge Volumes over Time. The upper-bound<br />

case simulated in the steady-state mass balance model may be too conservative (by providing<br />

larger concentration estimates) to accurately represent downstream concentrations of chemicals<br />

since effluent concentrations, treatment plant discharge volumes, and flow rates change over time.<br />

Therefore, the EPA will also use an empirical transport model originally developed by the USGS<br />

(Jobson, 1996) to simulate impacts from varying monthly discharge volumes over time. The<br />

50 Information on WWTF discharges in Pennsylvania can be found at https://www.paoilandgasreporting.state.pa.us/<br />

publicreports/Modules/Welcome/Welcome.aspx.<br />

51 Impacted watersheds may also have other sources of compounds of interest, possibly acid mine drainage and coal-fired<br />

utility boilers. This is discussed in more detail in Section 5.1, which also outlines work being done by the EPA to assess the<br />

contribution of hydraulic fracturing wastewater to contamination in surface water bodies.<br />

77


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

empirical approach is based on tracer studies performed around the United States since the early<br />

1970s (e.g., Nordin and Sabol (1974)). The empirical equations address two major difficulties in<br />

applying models to chemical transport scenarios: the inability to estimate travel times from crosssectional<br />

data and the reduction of concentration due to turbulent diffusion. The empirical equation<br />

approach gives an estimate of travel time and peak concentration so that the model does not need<br />

to be calibrated to tracer data.<br />

Hybrid Empirical-Numerical Model Estimates Impacts for River Networks. The original empirical<br />

approach was suited for a single river segment, or reach, of spatially uniform properties. The hybrid<br />

empirical-numerical model being developed by the EPA to expand the capabilities of the justdescribed<br />

Jobson technique will easily account for multiple reaches that can form branching river<br />

networks. Similar to all statistical relationships, the empirical equations do not always match tracer<br />

data exactly; therefore, the EPA is including the ability to perform Monte Carlo techniques in the<br />

software being developed. The EPA will confirm the accuracy of the hybrid model with tracer data<br />

that fall within the range of Jobson’s original set of inputs (taken from Nordin and Sabol (1974)) as<br />

well as later data from the Yellowstone River that provide a real-world test of this approach<br />

(McCarthy, 2009).<br />

The numerical portion of the hybrid model provides a direct and automatic comparison with the<br />

empirical equations. The method is based on a finite difference solution to the transport equation<br />

using recent developments in modeling to improve accuracy (Hairer et al., 1991; Leonard, 2002;<br />

Schiesser, 1991; Wallis, 2007). By including this numerical method, a hybrid empirical-numerical<br />

approach can be achieved. The empirical travel times from Jobson (1996) can be used to<br />

parameterize velocity in the numerical method. Dispersion coefficients can be derived from<br />

empirical data or a method developed by Deng et al. (2002). Using these approaches provides<br />

improved accuracy in the simulation results. The EPA will prepare a user’s guide to the model and<br />

make both the computer model and user’s guide widely available for duplicating the results<br />

prepared for this project and for more general use.<br />

For the generic simulations described above, effluent concentrations and discharge volumes will be<br />

modeled directly as variable inputs based on the effluent data evaluation (as discussed next in<br />

Section 4.2.4), while flow conditions will be modeled as low, medium, and high flow. Because the<br />

parameters describing transport are uncertain, statistical measures and Monte Carlo techniques<br />

will be used to generate probabilistic outputs from the model. To provide further assurance of the<br />

accuracy of the EPA hybrid model results, the Water Quality Simulation Package has been used to<br />

simulate tracer data and confirm the results (Ambrose et al., 1983; Ambrose and Wool, 2009;<br />

DiToro et al., 1981).<br />

4.2.4. Status and Preliminary Data<br />

The models described above are being used to determine potential impacts of treated wastewater<br />

discharges on downstream PWSs. Enough data have been identified to perform generic simulations<br />

for the steady-state mass balance simulations and hybrid empirical-numerical models with variable<br />

effluent concentration and plant discharge. For two WWTFs in Pennsylvania, USGS flow data have<br />

been compiled for segments of the rivers that reach downstream to drinking water intakes (50 to<br />

100 miles downstream) for the two locations. These data will be used to generate realistic model<br />

78


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

inputs to assess, in a generic sense, the potential impacts of discharges from realistic treatment<br />

plants.<br />

The EPA-developed hybrid empirical-numerical model has been favorably compared against a<br />

tracer experiment used by Jobson (1996) in developing the original empirical formulas. Calibration<br />

or other parameter adjustment was unnecessary for the hybrid model to produce accurate results.<br />

The EPA plans to compare the hybrid model to five more of the tracer experiments to cover the<br />

range of flow conditions used by Jobson (1996). Additionally, data from the more recent<br />

Yellowstone River experiment (McCarthy, 2009) are being prepared for testing the hybrid model.<br />

Similar comparisons of empirical to tracer experiments were performed by Reed and Stuckey<br />

(2002) for streams in the Susquehanna River Basin. The EPA Water Quality Simulation Package<br />

numerical model was set up to simulate the same tracer experiment performed for the hybrid<br />

model. Additional calibration is planned to refine the results from the Water Quality Simulation<br />

Package. After completing the evaluation of the hybrid model, the WWTF simulations will be<br />

completed.<br />

4.2.5. Next Steps<br />

A description of the EPA-developed empirical-numerical model and application of the empiricalnumerical<br />

and mass balance models to tracer experiments is being developed by EPA scientists and<br />

are expected to be submitted for publication in a peer-reviewed journal. The results from testing of<br />

the models and the analysis of the WWTF effluent data will be included in another peer-reviewed<br />

journal article.<br />

4.2.6. Quality Assurance Summary<br />

The initial QAPP for “Surface Water Transport of Hydraulic Fracturing-Derived Waste Water” was<br />

approved by the designated EPA QA Manager on September 8, 2011 (US EPA, 2012s). The QAPP<br />

was subsequently revised and approved on February 22, 2012.<br />

A TSA was conducted on March 1, 2012. The designated EPA QA Manager found the methods in use<br />

satisfactory and further recommendations for improving the QA process were unnecessary. An<br />

audit of data quality (ADQ) will be performed to verify that the quality requirements specified in<br />

the approved QAPP were met.<br />

79


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

4.3. Water Availability Modeling<br />

The EPA selected humid and semi-arid river basins as study areas for identifying potential impacts<br />

to drinking water resources from large volume water withdrawals (1 to 9 million gallons per well<br />

for the selected river basins) associated with hydraulic fracturing operations. This work is expected<br />

to address the research questions listed in Table 29.<br />

Table 29. Research questions addressed by modeling water withdrawals and availability in selected river basins.<br />

Water Cycle Stage<br />

Water acquisition<br />

Applicable Research Questions<br />

• How much water is used in hydraulic fracturing operations, and<br />

what are the sources of this water<br />

• How might water withdrawals affect short- and long-term water<br />

availability in an area with hydraulic fracturing<br />

• What are the possible impacts of water withdrawals for<br />

hydraulic fracturing operations on local water quality<br />

4.3.1. Project Introduction<br />

The volume of water needed in the hydraulic fracturing process for stimulation of unconventional<br />

oil and gas wells depends on the type of formation (e.g., coalbed, shale, or tight sands), the well<br />

construction (e.g., depth, length, vertical or directional drilling), and fracturing operations (e.g.,<br />

fracturing fluid properties and fracture job design). Water requirements for hydraulic fracturing of<br />

CBM range from 50,000 to 250,000 gallons per well (Holditch, 1993; Jeu et al., 1988; Palmer et al.,<br />

1991; Palmer et al., 1993), although much larger volumes of water are produced during the lifetime<br />

of a well in order to lower the water table and expose the coal seam (ALL Consulting, 2003; S.S.<br />

Papadopulos & Associates Inc., 2007a, b). The water usage for hydraulic fracturing in shale gas<br />

plays is significantly larger than CBM reservoirs—2 to 4 million gallons of water are typically<br />

needed per well (API, 2010; GWPC and ALL Consulting, 2009; Satterfield et al., 2008). The volume<br />

of water needed for well drilling is understood to be much less, from 60,000 gallons in the<br />

Fayetteville Shale to 1 million gallons in the Haynesville Shale (GWPC and ALL Consulting, 2009).<br />

Water-based mud systems used for drilling vertical or horizontal wells generally require that fresh<br />

water (non-potable, potable, or treated) be used as makeup fluid, although wells can also be drilled<br />

using compressed air and oil-based fluids.<br />

Water needed for hydraulic fracturing may come from multiple sources with varying quality.<br />

Sources may include raw surface and ground water, treated water from public water supplies, and<br />

water recycled from other purposes such as flowback and produced water from previous oil and<br />

gas operations or even acid mine drainage. The quality of water needed is dependent on the other<br />

chemicals in the fracturing fluid formulations, availability of water source, and the chemical and<br />

physical properties of the formation. The goal of this project is to investigate the water needs and<br />

sources to support hydraulic fracturing operations at the river basin and county spatial scales and<br />

to place this demand in the watershed context in terms of annual, seasonal, and monthly water<br />

availability.<br />

The EPA recognizes the unique circumstances of the geography and geology of every<br />

unconventional oil and gas resource and has chosen two study sites to initially explore and identify<br />

80


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

the potential differences related to water acquisition. The study areas includes two river basins: the<br />

Susquehanna River Basin (SRB), located in the eastern United States (humid climate) and overlying<br />

the Marcellus Shale gas reservoir (Figure 20), and the Upper Colorado River Basin (UCRB), located<br />

in the western United States (semi-arid climate) and overlying the Piceance structural basin and<br />

tight gas reservoir (Figure 21). The EPA is calibrating and testing watershed models for the study;<br />

the SRB and UCRB watershed models were previously calibrated and tested in the EPA<br />

investigation of future climate change impacts on watershed hydrology (the “20 watersheds study”)<br />

(Johnson et al., 2011).<br />

Figure 20. The Susquehanna River Basin, overlying a portion of the Marcellus Shale, is one of two study areas<br />

chosen for water availability modeling. Water acquisition for hydraulic fracturing will focus on Bradford and<br />

Susquehanna Counties in Pennsylvania. (GIS data obtained from ESRI, 2010a; US EIA, 2011e; US EPA, 2007.)<br />

81


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Figure 21. The Upper Colorado River Basin, overlying a portion of the Piceance Basin, is one of two river basins<br />

chosen for water availability modeling. Water acquisition for hydraulic fracturing will focus on Garfield and Mesa<br />

Counties in Colorado. (GIS data obtained from ESRI, 2010a; US EIA, 2011e; US EPA, 2007.)<br />

In both study areas, the river watershed and its subsurface basin include the river flows and<br />

reservoir and aquifer storages based on the hydrologic cycle, geography, geology, and water uses.<br />

The EPA’s goal is to explore future hypothetical scenarios of hydraulic fracturing use in the eastern<br />

and western study areas based on current understanding of hydraulic fracturing water acquisition<br />

and watershed hydrology. The EPA intends to characterize the significance, or insignificance, of<br />

hydraulic fracturing water use on future drinking water resources for the two study areas. The<br />

research will involve detailed representation of water acquisition supporting hydraulic fracturing<br />

in the Bradford County and Susquehanna County area in Pennsylvania and in the Garfield County<br />

and Mesa County areas of Colorado. These areas have concentrated hydraulic fracturing activity, as<br />

discussed below.<br />

4.3.1.1. Susquehanna River Basin<br />

Geography, Hydrology, and Climate. The SRB has over 32,000 miles of waterways, drains 27,510<br />

square miles, and covers half of Pennsylvania and portions of New York and Maryland (Figure 20)<br />

(SRBC, 2006). On average, the SRB contributes 18 million gallons of water every minute (25,920<br />

million gallons per day, or MGD) to the Chesapeake Bay (SRBC, 2006). The humid climate of the<br />

region experiences long-term average precipitation of 37 to 43 inches per year (McGonigal, 2005).<br />

82


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Oil and Gas Resources and Activity. Large portions of the SRB watershed are underlain by the<br />

Marcellus Shale formation, which is rich in natural gas. Estimates of recoverable and undiscovered<br />

natural gas from this formation range from 42 to 144 trillion cubic feet (Coleman et al., 2011) and<br />

production well development estimates for the next two decades range as high as 60,000 total wells<br />

drilled by 2030 (Johnson et al., 2010). The Pennsylvania Department of Environmental Protection<br />

reports that the number of drilled wells in the Marcellus Shale has been increasing rapidly. In 2007,<br />

only 27 Marcellus Shale wells were drilled in the state; in 2010 the number of wells drilled was<br />

1,386. Data extracted from FracFocus 52 indicate that the total vertical depth of wells in Bradford<br />

and Susquehanna Counties is between 5,000 and 8,500 feet (mean of 6,360 feet) below ground<br />

surface, which implies that this depth range is the target production zone for the Marcellus Shale.<br />

Water Use. The SRB supports a population of over 4.2 million people. Table 30 lists the estimated<br />

water use for the SRB and Bradford and Susquehanna Counties. The Susquehanna River Basin<br />

Commission estimates consumptive water use in five major categories, with PWSs consuming the<br />

greatest volume of water per day (325 MGD) followed by thermoelectric energy production (190<br />

MGD) (Richenderfer, 2011). The greatest water withdrawals per day in Bradford and Susquehanna<br />

Counties are for drinking water (8.25 MGD for combined public and domestic use) and self-supplied<br />

industrial uses (4.59 MGD).<br />

Table 30. Water withdrawals for use in the Susquehanna River Basin (Richenderfer, 2011) and Bradford and<br />

Susquehanna Counties, Pennsylvania (Kenny et al., 2009).<br />

Use<br />

Public supply<br />

Self-supplied domestic<br />

Irrigation (crop)<br />

Irrigation (golf courses)<br />

Self-supplied industrial<br />

Livestock<br />

Thermoelectric<br />

Mining<br />

Other<br />

Water Withdrawals (million gallons per day)<br />

Susquehanna River Basin<br />

Not reported<br />

Bradford and Susquehanna<br />

Counties, Pennsylvania<br />

325 4.59<br />

Not reported 3.66<br />

Not reported<br />

22.0<br />

Not reported<br />

190<br />

(energy production, non-gas)<br />

10.0<br />

50.0<br />

(recreation)<br />

0.110<br />

0.060<br />

4.59<br />

3.41<br />

0.00<br />

0.10<br />

Not reported<br />

Figure 22 displays the geographic distribution of PWSs in the SRB. 53<br />

52 See Section 3.5 for additional information on the FracFocus data extraction and analysis research project.<br />

53 The location and type of drinking water supply is significant when represented in watershed hydrology models. The<br />

extraction of surface water is removed from the watershed model subbasin from its main river reach. The extraction of<br />

ground water is removed from the model subbasin from its ground water storage.<br />

83


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Figure 22. Public water systems in the Susquehanna River Basin (US EPA, 2011j). The legend symbol size for public<br />

water systems is proportional to the number of people served by the systems. For example, the smallest circle<br />

represents water systems serving 25 to 100 people and the largest circle represents systems serving over 100,000<br />

people.<br />

The Susquehanna River Basin Commission reports that the oil and gas industry consumed over 1.6<br />

billion gallons of water for well drilling and hydraulic fracturing in the entire SRB from July 1, 2008,<br />

to February 14, 2011. If averaged over the entire time, this is roughly 1.7 MGD. This amount of<br />

water was used for approximately 1,800 gas production wells with about 550 wells hydraulically<br />

fractured by the end of 2010 (Richenderfer, 2011). The majority (65%) of the water came from<br />

direct surface water withdrawals, with smaller fractions from PWSs (35%) and ground water (very<br />

small). The average total volume of fluid used per well was 4.2 million gallons, with about 10% of<br />

the volume as treated flowback and 90% fresh water (Richenderfer, 2011). The average recovery of<br />

fluids was reported to be 8% to 12% of the injected volume within the first 30 days (Richenderfer,<br />

2011).<br />

84


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Water use reported in FracFocus for Bradford and Susquehanna Counties ranges between 2 and 9<br />

million gallons per well (median of 4.7 million gallons per well; (GWPC, 2012a)), consistent with<br />

data reported by the Susquehanna River Basin Commission. 54 In this part of the SRB, the wells are<br />

almost exclusively horizontal and producing from the Marcellus Shale. The operators are blending<br />

treated produced water into hydraulic fracturing fluids (Rossenfoss, 2011).<br />

4.3.1.2. Upper Colorado River Basin<br />

Geography, Hydrology, and Climate. The UCRB drains an area of 17,800 square miles and is<br />

characterized by high mountains in the east and plateaus and valleys in the west. The average<br />

discharge of the Colorado River near the Colorado-Utah state line is about 2.8 million gallons per<br />

minute (about 4,000 MGD) (Coleman et al., 2011). Precipitation ranges from 40 inches per year or<br />

more in the eastern part of the basin to less than 10 inches per year in the western part of the basin<br />

(Spahr et al., 2000).<br />

Oil and Gas Resources and Activity. The UCRB has a long history of oil, gas, and coal exploration. The<br />

Piceance Basin is a source of unconventional natural gas and oil shale. The basin was originally<br />

exploited for its coal resources, and the associated CBM production peaked around 1992 (S.S.<br />

Papadopulos & Associates Inc., 2007a). The Upper Cretaceous Williams Fork Formation, a thick<br />

section of shale, sandstone, and coal, has been recognized as a significant source of gas since 2004<br />

(Kuuskraa and Ammer, 2004). The wells producing gas from the Williams Fork are either vertically<br />

or directionally (“S”-shaped wells) drilled rather than horizontal. While the deeper Mancos Shale is<br />

considered a major resource for shale gas (Brathwaite, 2009), it must be exploited with horizontal<br />

drilling methods, and the economics are such that only prospecting wells are being drilled at this<br />

time (personal communication, Jonathan Shireman, Shaw Environmental & Infrastructure, May 1,<br />

2012). Estimated reserves in coalbeds and unconventional tight gas reservoirs are nearly 84 trillion<br />

cubic feet (Tyler and McMurry, 1995).<br />

Gas production activities occur in the following counties within the UCRB: Delta, Eagle, Garfield,<br />

Grand, Gunnison, Hinsdale, Mesa, Montrose, Ouray, Pitkin, Routt, Saguache, and Summit (COGCC,<br />

2012b). Table 31 indicates that the greatest drilling activity has been in Garfield and Mesa Counties<br />

(Figure 21), where well completions increased steadily from 2000 (212 wells) to 2008 (2,725<br />

wells), then dropped slightly to 1,160 wells in 2010 (COGCC, 2012b). The total vertical depth of<br />

wells in Garfield County and Mesa County as reported in FracFocus implies that the location of the<br />

target production zone(s) lies between 6,000 and 13,000 feet (mean of 8,000 feet) below ground<br />

surface.<br />

54 More information on FracFocus is available in Section 3.5.<br />

85


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table 31. Well completions for select counties in Colorado within the Upper Colorado River Basin watershed<br />

(COGCC, 2012b).<br />

County<br />

2000<br />

2001<br />

Annual Well Completions from 2000 to 2010<br />

2002 2003 2004 2005 2006 2007 2008<br />

2009<br />

2010<br />

Delta<br />

8 5 8 3 2<br />

4<br />

Garfield<br />

Gunnison<br />

207<br />

244<br />

287 507 679 892 1269 1689 2255<br />

2 3 2 1 11 8 2<br />

1050<br />

4<br />

1139<br />

2<br />

Mesa<br />

5<br />

21<br />

26 18 53 203 336 501 470<br />

43<br />

21<br />

Montrose<br />

4<br />

2 2 3 4<br />

1<br />

Routt<br />

10<br />

21<br />

8 5 2<br />

4<br />

1<br />

Water Use. The UCRB supports a population of over 275,000 people. Table 32 lists the estimated<br />

water use for the UCRB and Garfield and Mesa Counties in Colorado. According to the USGS, the<br />

total water use in 2005 in the UCRB and Garfield and Mesa Counties was dominated by irrigation<br />

(1702 and 1200 MGD, respectively), followed by public and domestic water supply (60.4 and 29.6<br />

MGD), and thermoelectric energy production (44 MGD) (Ivahnenko and Flynn, 2010; Kenny et al.,<br />

2009).<br />

Table 32. Water withdrawals for use in the Upper Colorado River Basin (Ivahnenko and Flynn, 2010) and Garfield<br />

and Mesa Counties in Colorado (Kenny et al., 2009).<br />

Use<br />

Water Withdrawals (million gallons per day)<br />

Upper Colorado River Basin<br />

Garfield and Mesa Counties, Colorado<br />

Public supply 58.6 29.2<br />

Self-supplied domestic 1.81 1.35<br />

Irrigation (crop) 1702 1200<br />

Irrigation (golf courses) 8.00 3.50<br />

Self-supplied industrial 2.71 1.05<br />

Livestock 0.870 0.840<br />

Thermoelectric<br />

43.9<br />

(non-consumptive)<br />

43.9<br />

(non-consumptive)<br />

Mining 0.390 0.280<br />

Other<br />

Not reported<br />

1.88<br />

(aquaculture)<br />

Figure 23 displays the distribution of public water systems in the basin. Interbasin water transfers,<br />

mining, urbanization, and agriculture are the principal human activities that potentially impact<br />

water quantity in the UCRB.<br />

86


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Figure 23. Public water systems in the Upper Colorado River Basin (US EPA, 2011j). The legend symbol size for<br />

public water systems is proportional to the number of people served by the systems. For example, the smallest circle<br />

represents water systems serving 25 to 100 people and the largest circle represents systems serving over 70,000<br />

people.<br />

The State of Colorado estimates that total annual statewide water demand for hydraulic fracturing<br />

associated with oil and gas wells increased from 4.5 billion gallons in 2010 to almost 4.9 billion<br />

gallons in 2011 (12.3 MGD in 2010 to almost 13.4 MGD in 2011), which parallels the increasing<br />

number of wells spudded, as shown in Table 33 (COGCC, 2012a). The amount of water demand was<br />

determined using the number of wells spudded (horizontal and vertical) multiplied by an average<br />

amount of water required for hydraulic fracturing per well type based on data reported in 2011.<br />

COGCC (2012a) estimates the average water use per well at about 1.6 million gallons in 2010 and<br />

2011.<br />

87


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table 33. Estimated total annual water demand for oil and gas wells in Colorado that were hydraulically fractured in<br />

2010 and 2011 (COGCC, 2012a). Data for vertical and horizontal wells are not differentiated in the estimates and well<br />

spud dates.<br />

Category<br />

Year<br />

2010 2011<br />

Wells spudded 2,753 2,975<br />

Estimated annual water demand<br />

(million gallons)<br />

Estimated water use per well<br />

(million gallons)<br />

4,531 4,857<br />

1.65 1.63<br />

Data extracted from FracFocus for Garfield and Mesa Counties shows water use per well between 1<br />

and 9 million gallons (median 1.3 million gallons), which is consistent with the Colorado Oil and Gas<br />

Compact Commission data (COGCC, 2012a; GWPC, 2012a). In this part of the Piceance Basin (Figure<br />

21), the majority of wells are vertically drilled and producing gas from the Williams Fork tight<br />

sandstones. Based on conversations with Berry Petroleum, Williams Production, Encana Oil and<br />

Gas, and the Colorado Field Office of the US Bureau of Land Management, the water used to fracture<br />

wells in this area is entirely recycled formation water that is recovered during production<br />

operations. Fresh water is used only for drilling mud, cementing the well casing, hydrostatic testing,<br />

and dust abatement and is estimated to be about 251,000 gallons per well (US FWS, 2008).<br />

4.3.2. Research Approach<br />

Watershed Models. In order to assess the impact of hydraulic fracturing water withdrawals on<br />

drinking water availability at watershed and county spatial scales as well as annual, seasonal,<br />

monthly, and daily time scales, the EPA is developing separate hydrologic watershed models for the<br />

SRB and UCRB. The models are based in part on the calibrated and verified watershed models<br />

(hereafter called the “foundation” models) of the EPA Global Change Research Program (Johnson et<br />

al., 2011), namely the Hydrologic Simulation Program FORTRAN (HSPF) 55 and the Soil and Water<br />

Assessment Tool (SWAT). 56 Both HSPF and SWAT are physically based, semi-distributed watershed<br />

models that compute changes in water storage and fluxes within drainage areas and water bodies<br />

over time. Each model can simulate the effect of water withdrawals or flow regulation on modeled<br />

stream or river flows. Key inputs for the models include meteorological data, land use data, and<br />

time series data representing water withdrawals. The models give comparable performance at the<br />

scale of investigation (Johnson et al., 2011).<br />

Modeling of the SRB will be completed using the calibrated and tested HSPF. Since its initial<br />

development nearly 20 years ago, HSPF has been applied around the world; it is jointly sponsored<br />

by the EPA and the USGS, and has extensive documentation and references (Donigian Jr., 2005;<br />

Donigian Jr. et al., 2011). The choice of HSPF in the SRB, a subwatershed within the larger<br />

55 More information on the HSPF model including self-executable file, is available at http://www.epa.gov/ceampubl/<br />

swater/hspf/.<br />

56 More information on the SWAT model including self-executable file, is available at http://swat.tamu.edu/<br />

software/swat-model/.<br />

88


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Chesapeake Bay watershed, allows benchmarking to the peer-reviewed and community-accepted<br />

Chesapeake Bay Program watershed model. 57<br />

Modeling of the UCRB will be completed using the calibrated and tested SWAT. The SWAT is a<br />

continuation of over 30 years of modeling efforts conducted by the US Department of Agriculture’s<br />

Agricultural Research Service and has extensive peer review (Gassman et al., 2007). SWAT is an<br />

appropriate choice in the less data-rich UCRB, where hydrological response units can be<br />

parameterized based on publicly available GIS maps of land use, topography, and soils.<br />

The SRB and UCRB models will build on the “foundation” models and be updated to represent<br />

baseline and current watershed conditions. The baseline model will add reservoirs and major<br />

consumptive water uses for watershed conditions of the year 2000 for the SRB and 2005 for the<br />

UCRB. The baseline year predates the significant expansion of hydraulic fracturing in the basin<br />

(2007 for SRB, 2008 for UCRB) and corresponds with the USGS’ water use reports (every five years<br />

since 1950) and the National Land Cover Dataset (Homer et al., 2007). The baseline models will<br />

represent the USGS’s major water use categories, including the consumptive component of both<br />

PWS and domestic water use, and the other major water use categories (irrigation, livestock,<br />

industrial, mining, thermoelectric power). The snapshot of each watershed in the year 2010 will be<br />

the current model representation in both basins. The current models will include all water use<br />

categories from the baseline model plus hydraulic fracturing water withdrawals and refine the<br />

representation of PWS and hydraulic fracturing in county-scale focus areas—Garfield/Mesa<br />

Counties in Colorado and Bradford/Susquehanna Counties in Pennsylvania.<br />

The foundation, baseline, and current watershed models will be exposed to the historical<br />

meteorology (precipitation, temperature) from National Weather Service gauges located within<br />

each watershed. The calibration and validation of the foundation, baseline, and current models will<br />

be checked by comparing goodness-of-fit statistics and through expert judgment of comparisons of<br />

observed and modeled stream discharges.<br />

Key characteristics of model configuration include:<br />

• Land use will be based on the 2001 National Land Cover Dataset (Homer et al., 2007).<br />

Land use data are used for segmenting the basin land area into multiple hydrologic<br />

response units, each with unique rainfall/runoff response properties. For the SWAT<br />

model, soil and slope data will also be used for defining unique hydrologic response units.<br />

• Each basin will be segmented into multiple subwatersheds at the 10-digit hydrologic unit<br />

scale. 58<br />

57 More information on the Chesapeake Bay Program watershed model is available at http://www.chesapeakebay.net/<br />

about/programs/modeling/53/.<br />

58 Hydrologic units refer to the Watershed Boundary Dataset developed through a coordinated effort by the USGS, the US<br />

Department of Agriculture, and the EPA. The intent of defining hydrologic units for the Watershed Boundary Dataset is to<br />

establish a baseline drainage boundary framework, accounting for all land and surface areas. Several levels of watershed<br />

are defined based on size. A 10-digit hydrologic unit is a level 5 watershed of average size 227 square miles (USDA, 2012).<br />

89


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

• Observed meteorological data for water years 1972 to 2004 for SRB and 1973 to 2003 for<br />

UCRB will be applied to assess water availability over a range of weather conditions.<br />

• The effect of reservoirs on downstream flows will be simulated using reservoir<br />

dimensions/operation data from circa 2000 from the Chesapeake Bay Program watershed<br />

model (Phase 5.3; (US EPA, 2010a)).<br />

• Point source dischargers with NPDES-permitted flow rates of at least 1 MGD will be<br />

represented as sources of water on the appropriate stream reaches.<br />

• Surface water withdrawals will be simulated for three unique water use categories:<br />

hydraulic fracturing water use, PWSs, and other. For the “other” category, the magnitude<br />

of withdrawals from modeled stream reaches will be based on water use estimates<br />

developed by the USGS (year 2000 for SRB; year 2005 for UCRB). 59<br />

Modeling Future Scenarios. The modeling effort will also simulate a snapshot of heightened annual<br />

hydraulic fracturing relative to the baseline and current condition models at levels that could<br />

feasibly occur over the next 30 years, based on recent drilling trends and future projections of<br />

natural gas production (US EIA, 2012; US EPA, 2012w). Because projections of future conditions are<br />

inherently uncertain, three separate scenarios will be simulated: business-as-usual, energy plus,<br />

and green technology. The scenarios assume distinct levels of natural gas drilling and hydraulic<br />

fracturing freshwater use and, therefore, apply distinct hydraulic fracturing water withdrawal time<br />

series to modeled stream reaches. Further, significant population growth is projected in<br />

Garfield/Mesa Counties, Colorado, over the next 30 years (US EPA, 2010c), where natural gas<br />

extraction in the UCRB has recently been concentrated. Therefore, the UCRB future scenarios also<br />

consider a potential increase in PWS surface withdrawals in the basin. The balance between surface<br />

water availability and demand depicted in each scenario’s annual snapshot of water use will be<br />

assessed across a range of weather conditions (i.e., drought, dry, wet, and very wet years based on<br />

the historical record). A description of each scenario, and the methods used for scenario<br />

development, are provided below and in Tables 34 and 35.<br />

59 The USGS water use estimates can be found at http://water.usgs.gov/watuse/.<br />

90


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table 34. Data and assumptions for future watershed availability and use scenarios modeled for the Susquehanna<br />

River Basin. Current practices for water acquisition and disposal are tracked by the Susquehanna River Basin<br />

Commission (SRBC).<br />

Model Assumptions<br />

Hydraulic fracturing<br />

well deployment<br />

Hydraulic fracturing<br />

water management<br />

practices<br />

* US EPA, 2012w; USGS, 2011c<br />

†<br />

SRBC, 2012<br />

Business as Usual<br />

Current well inventory<br />

and future deployment<br />

schedules and playlevel<br />

development<br />

projections*<br />

Current practices for<br />

water acquisition,<br />

production and disposal<br />

tracked by SRBC †<br />

Future Scenarios<br />

Energy Plus<br />

Maximum projected<br />

development of gas<br />

reserves*<br />

Current practices for<br />

water acquisition,<br />

production and disposal<br />

tracked by SRBC †<br />

Green Technology<br />

Current well inventory<br />

and future deployment<br />

schedules and playlevel<br />

development<br />

projections*<br />

Increased recycling of<br />

produced water for<br />

hydraulic fracturing †<br />

Table 35. Data and assumptions for future watershed availability and use scenarios modeled for the Upper Colorado<br />

River Basin.<br />

Model Assumptions<br />

Future Scenarios<br />

Business as Usual Energy Plus*<br />

Current well inventory<br />

and future deployment Maximum projected<br />

Hydraulic fracturing<br />

schedules and playlevel<br />

development reserves †<br />

development of gas<br />

well deployment<br />

projections †<br />

Current practices for Current practices for<br />

Hydraulic fracturing<br />

water acquisition, water acquisition,<br />

water management<br />

production and disposal production and disposal<br />

practices<br />

estimated for UCRB § estimated for UCRB §<br />

* Reflects 2040 population increase (US EPA, 2010c) and corresponding change in PWS demand.<br />

†<br />

US EIA, 2011b, 2012; US EPA, 2012w; USGS, 2003<br />

§<br />

US FWS, 2008<br />

Green Technology*<br />

Maximum projected<br />

development of gas<br />

reserves †<br />

Increased recycling of<br />

produced water for<br />

drilling §<br />

Future drilling patterns in the SRB and UCRB are assessed from National Energy Modeling System<br />

(NEMS) regional projections of the number of wells drilled annually from 2011to 2040 in shale gas<br />

(SRB) and tight gas (UCRB) plays (US EIA, 2012; US EPA, 2012w). Based on analysis of NEMS well<br />

projections and undiscovered resources in the Marcellus Shale (Coleman et al., 2011), peak annual<br />

drilling in the SRB could exceed the recent high in 2011 by as much as 50%. In the UCRB, analysis of<br />

NEMS well projections and undiscovered tight gas resources in the Piceance Basin (USGS, 2003)<br />

suggest that the 2008 peak level of drilling in the basin could be repeated in the late 2030s, when a<br />

growing population would exert a higher demand for freshwater. The future scenarios will<br />

incorporate these projections, with high-end estimates of the number of wells drilled/fractured<br />

applied in the energy plus scenario.<br />

The volume of surface water required for drilling and hydraulic fracturing varies according to local<br />

geology, well characteristics, and the amount of recycled water available for injection. In the SRB,<br />

2008 to 2011 water use data (SRBC, 2012) show that, on average, 13% of total water injected for<br />

91


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

hydraulic fracturing is composed of recycled produced water or wastewater. Per well surface water<br />

use in the SRB business as usual and energy plus scenarios will therefore be established as 87% of<br />

the 4 million gallons of water used for hydraulic fracturing, or 3.5 million gallons. The SRB green<br />

technology scenario reflects a condition of increased water recycling, where the 90 th percentile of<br />

current recycled water amount (29%) becomes the average. Per well surface water use in the SRB<br />

green technology scenario will therefore be established as 71% of the 4 million gallons of water<br />

used for hydraulic fracturing, or 2.8 million gallons.<br />

In the UCRB, 100% recycled water use is typical for hydraulic fracturing of tight sandstones<br />

(personal communication, Jonathan Shireman, Shaw Environmental & Infrastructure, May 7, 2012).<br />

Surface water is acquired for well drilling and cementing (0.18 million gallons), dust abatement<br />

(0.03 million gallons), and hydrostatic testing (0.04 million gallons) only (US FWS, 2008). Per well<br />

surface water use in the UCRB business as usual and energy plus scenarios will therefore be 0.25<br />

million gallons. For the UCRB green technology scenario, surface water will be assumed to be<br />

acquired for well drilling and cementing only (0.18 million gallons per well).<br />

Following the development of water withdrawal datasets for each scenario, model output will be<br />

reviewed to assess the impacts of water acquisition for hydraulic fracturing on drinking water<br />

supplies by evaluating annual and long-term streamflow and water demand, and identifying shortterm<br />

periods (daily to monthly) in which water demand exceeds streamflow. Since many public<br />

water supplies originate from ground water sources, simulated ground water recharge will also be<br />

computed. Results will be compared among the three scenarios to identify noteworthy differences<br />

and their implications for future management of hydraulic fracturing-related water withdrawals.<br />

4.3.3. Status and Preliminary Data<br />

Existing water use information for hydraulic fracturing has been collected from the Susquehanna<br />

River Basin Commission and the Colorado Oil and Gas Compact Commission by Shaw<br />

Environmental Technologies. The data underwent a QA review before submission to the modeling<br />

teams of The Cadmus Group, Inc. The models are being calibrated and validated. The future<br />

scenarios are being designed, with model simulations to follow. Work is underway and will be<br />

published in peer-reviewed journals when completed.<br />

4.3.4. Quality Assurance Summary<br />

The QAPP, “Modeling the Impact of Hydraulic Fracturing on Water Resources Based on Water<br />

Acquisition Scenarios (Version 1.0),” contracted through The Cadmus Group, Inc., was accepted on<br />

February 8, 2012 (Cadmus Group Inc., 2012a). A technical directive/contract modification dated<br />

April 25, 2012, modifies the scope of the project but not the procedures. Additionally, there is a<br />

pending QAPP revision that adapts the scope to the contract modification.<br />

A TSA of The Cadmus Group, Inc., contract was performed by the designated EPA QA Manager on<br />

June 14, 2012. The methods in use were found to be satisfactory and further recommendations for<br />

improving the QA process were unnecessary. Work performed and scheduled to be performed was<br />

within the scope of the project.<br />

92


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

The interim progress report “Development and Evaluation of Baseline and Current Conditions for<br />

the Susquehanna River Basin,” received on June 19, 2012, was found to be concise but detailed<br />

enough to meet the QA requirements, as expressed in the QAPP, its revision, and the contract<br />

modification/technical directive. The same was true for the interim progress report “Impact of<br />

Water Use and Hydro-Fracking on the Hydrology of the Upper Colorado River Basin,” submitted on<br />

July 2, 2012.<br />

93


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

5. Laboratory Studies<br />

The laboratory studies are targeted research projects designed to improve understanding of the<br />

ultimate fate and transport of selected chemicals, which may be components of hydraulic fracturing<br />

fluids or naturally occurring substances released from the subsurface during hydraulic fracturing.<br />

This chapter includes progress reports for the following projects:<br />

5.1. Source Apportionment Studies........................................................................................................................ 94<br />

Identification and quantification of the source(s) of high bromide and chloride concentrations<br />

at public water supply intakes downstream from wastewater treatment plants discharging<br />

treated hydraulic fracturing wastewater to surface waters<br />

5.2. Wastewater Treatability Studies.................................................................................................................. 101<br />

Assessment of the efficacy of common wastewater treatment processes on removing selected<br />

chemicals found in hydraulic fracturing wastewater<br />

5.3. Brominated Disinfection Byproduct Precursor Studies ..................................................................... 107<br />

Assessment of the ability of bromide and brominated compounds present in hydraulic<br />

fracturing wastewater to form brominated disinfection byproducts (Br-DBPs) during drinking<br />

water treatment processes<br />

5.4. Analytical Method Development .................................................................................................................. 112<br />

Development of analytical methods for selected chemicals found in hydraulic fracturing fluids<br />

or wastewater<br />

5.1. Source Apportionment Studies<br />

5.1.1. Relationship to the Study<br />

The EPA is combining data collected from samples of wastewater treatment facility discharges and<br />

receiving waters with existing modeling programs to identify the proportion of hydraulic fracturing<br />

wastewater that may be contributing to contamination at downstream public water system intakes.<br />

This work has been designed to help inform the answer to the research question listed in Table 36.<br />

Table 36. Secondary research questions addressed by the source apportionment research project.<br />

Water Cycle Stage<br />

Wastewater treatment and<br />

waste disposal<br />

Applicable Research Questions<br />

What are the potential impacts from surface water disposal of treated<br />

hydraulic fracturing wastewater on drinking water treatment<br />

facilities<br />

5.1.2. Project Introduction<br />

The large national increase in hydraulic fracturing activity has generated large volumes of hydraulic<br />

fracturing wastewater for treatment and disposal or recycling. In some cases, states have allowed<br />

hydraulic fracturing wastewater to be treated by WWTFs with subsequent discharge to rivers. Most<br />

WWTFs are designed to filter and flocculate solids, as well as consume biodegradable organic<br />

species associated with human and some commercial waste. Very few facilities are designed to<br />

94


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

manage the organic and inorganic chemical compounds contained in hydraulic fracturing<br />

wastewater.<br />

Public water supply intakes may be located in river systems downstream from WWTFs and a<br />

variety of other industrial and urban discharges, and it is critical to evaluate sources of<br />

contamination at those drinking water intakes. Elevated bromide and chloride concentrations are<br />

of particular concern in drinking water sources due to the propensity of bromides to react with<br />

organic compounds to produce THMs and other DBPs during drinking water treatment processes<br />

(Plewa and Wagner, 2009). High TDS levels—including bromide and chloride—have been detected<br />

in the Monongahela River in 2008 and the Youghiogheny River in 2010 (Lee, 2011; Ziemkiewicz,<br />

2011). The source and effects of these elevated concentrations remains unclear.<br />

This project’s overall goal is to establish an approach whereby surface water samples may be<br />

evaluated to determine the extent to which hydraulic fracturing wastewaters (treated or untreated)<br />

may be present, and to distinguish whether any elevated bromide and chloride in those samples<br />

may be due to hydraulic fracturing or other activities. To accomplish this goal, the EPA is: (1)<br />

quantifying the inorganic chemical composition of discharges in two Pennsylvania river systems<br />

from WWTFs that accept and treat flowback and produced water, coal-fired utility boilers, acid<br />

mine drainage, stormwater runoff of roadway deicing material, and other industrial sources; (2)<br />

investigating the impacts of the discharges by simultaneously collecting multiple upstream and<br />

downstream samples to evaluate transport and dispersion of inorganic species; and (3) estimating<br />

the impact of these discharges on downstream bromide and chloride levels at PWS intakes using<br />

mathematical models.<br />

5.1.3. Research Approach<br />

The “Quality Assurance Project Plan for Hydraulic Fracturing Wastewater Source Apportionment”<br />

provides a detailed description of the research approach (US EPA, 2012q). Briefly, water samples<br />

are being collected at five locations on two river systems; each river has an existing WWTF that is<br />

currently accepting hydraulic fracturing wastewater for treatment. Source profiles for significant<br />

sources such as hydraulic fracturing wastewater, WWTF effluent, coal-fired utility boiler<br />

discharge, acid mine drainage, and stormwater runoff from roadway deicing will be developed<br />

from samples collected from these sources during the study. Computer models will then be used<br />

to compare data from these river systems to chemical and isotopic composition profiles obtained<br />

from potential sources.<br />

Three two-week intensive sampling events were conducted to assess river conditions under<br />

different flow regimes: spring, summer, and fall 2012. As shown in Table 37, the amount of water in<br />

the river has historically been highest in the spring, resulting in the dilution of pollutants, and the<br />

summer and fall seasons typically have decreased stream flow, which may result in elevated<br />

concentrations due to less dilution (USGS, 2011a, b). USGS gauging stations near the WWTFs will be<br />

used to measure the flow rate during the three sampling periods.<br />

95


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table 37. Historical average of monthly mean river flow and range of monthly means from 2006 through 2011 for two<br />

rivers in Pennsylvania where the EPA collects samples for source apportionment research (USGS, 2011a, b).<br />

Month<br />

Average of Monthly Mean River Flow Range of Monthly Means from 2006<br />

from 2006 Through 2011<br />

Through 2011<br />

(cubic feet per second)<br />

(cubic feet per second)<br />

Allegheny River Blacklick Creek Allegheny River Blacklick Creek<br />

May 12,100 357 7,330–28,010 220.2–479.7<br />

July 5,740 134 2,164–10,840 65.8–198.2<br />

September 4,940 174 2,873–13,560 48.8–520.0<br />

During each sampling event, automatic water samplers (Teledyne Isco, model 6712) at each site<br />

collect two samples daily—morning and afternoon—based on the PWS and WWTF operations<br />

schedule. The samples are stored in the sampler for one to four days, depending on the site visit<br />

schedule. Each river is sampled in five locations, as shown in Table 38. The first sampling device<br />

downstream of the WWTF is far enough downstream to allow for adequate mixing of the WWTF<br />

effluent and river water. The second downstream sampling device is between the first<br />

downstream sampling location and the closest PWS intake. The locations of the samplers<br />

downstream of the WWTF also take into account the presence of other significant sources, such<br />

as coal-fired utility boiler and acid mine drainage discharges, and allow for the evaluation of their<br />

impacts.<br />

Table 38. Distance between sampling sites and wastewater treatment facilities on two rivers where the EPA collects<br />

samples for source apportionment research<br />

Site<br />

.<br />

Distance Between Sampling Sites (kilometers)<br />

Allegheny River<br />

Blacklick Creek<br />

Site 1 (upstream) -1.6 -1.2<br />

Site 2 (wastewater treatment facility) 0 0<br />

Site 3 (downstream) 12.2 2.7<br />

Site 4 (downstream) 44.1 43.1<br />

Site 5 (public water system intake) 52.3 88.6<br />

5.1.3.1. Sample Analyses<br />

The EPA will analyze the river samples and effluent samples according to existing EPA methods for<br />

the suite of elements and ions listed in Table 39. Inorganic ions (anions and cations) are being<br />

determined by ion chromatography. Inorganic elements are being determined using a combination<br />

of inductively coupled plasma optical emission spectroscopy for high-concentration elements and<br />

high-resolution magnetic sector field inductively coupled plasma mass spectrometry for low<br />

concentration elements. Additionally, the characteristic strontium (Sr) ratios ( 87 Sr/ 86 Sr; 0.7101–<br />

0.7121) in Marcellus Shale brines are extremely sensitive tracers, and elevated concentrations of<br />

readily water soluble strontium are present in the hydraulic fracturing wastewaters (Chapman et<br />

al., 2012). Isotope analyses for 87 Sr/ 86 Sr are being conducted on a subset (~20%) of samples by<br />

thermal ionization mass spectrometry to corroborate source apportionment modeling results.<br />

96


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table 39. Inorganic analyses and respective instrumentation planned for source apportionment research. The EPA<br />

will analyze samples from two rivers and effluent discharged from wastewater treatment facilities located on each<br />

river. Instruments used for analysis include high-resolution magnetic sector field inductively coupled plasma mass<br />

spectrometry (HR-ICP-MS), ion chromatography (IC), inductively coupled plasma optical emission spectroscopy<br />

(ICP-OES), and thermal ionization mass spectroscopy (TIMS).<br />

Element Instrument Used<br />

Element Instrument Used<br />

Ag*<br />

HR-ICP-MS<br />

Sb*<br />

HR-ICP-MS<br />

Al*<br />

ICP-OES<br />

Sc<br />

HR-ICP-MS<br />

As*<br />

HR-ICP-MS<br />

Se*<br />

HR-ICP-MS<br />

B* ICP-OES<br />

Si<br />

ICP-OES<br />

Ba*<br />

ICP-OES<br />

Sm<br />

HR-ICP-MS<br />

Be*<br />

HR-ICP-MS<br />

Sn<br />

HR-ICP-MS<br />

Bi<br />

HR-ICP-MS<br />

Sr*<br />

HR-ICP-MS<br />

Ca*<br />

ICP-OES<br />

Tb<br />

HR-ICP-MS<br />

Cd*<br />

HR-ICP-MS<br />

Th<br />

HR-ICP-MS<br />

Ce<br />

HR-ICP-MS<br />

Ti*<br />

ICP-OES<br />

Co*<br />

HR-ICP-MS<br />

Tl*<br />

HR-ICP-MS<br />

Cr*<br />

HR-ICP-MS<br />

U<br />

HR-ICP-MS<br />

Cs*<br />

HR-ICP-MS<br />

V* HR-ICP-MS<br />

Cu* ICP-OES, HR-ICP-MS<br />

W<br />

HR-ICP-MS<br />

Fe* ICP-OES, HR-ICP-MS<br />

Y<br />

HR-ICP-MS<br />

Gd<br />

HR-ICP-MS<br />

Zn*<br />

ICP-OES<br />

Ge<br />

HR-ICP-MS<br />

Isotope Ratio Instrument Used<br />

K* ICP-OES<br />

87 Sr/ 86 Sr* TIMS<br />

La<br />

HR-ICP-MS<br />

Ion<br />

Instrument Used<br />

Li*<br />

ICP-OES<br />

Ca 2+ *<br />

IC<br />

Mg*<br />

ICP-OES<br />

K + *<br />

IC<br />

Mn* ICP-OES, HR-ICP-MS<br />

Li + *<br />

IC<br />

Mo*<br />

HR-ICP-MS<br />

Mg 2+ *<br />

IC<br />

Na*<br />

ICP-OES<br />

+<br />

NH 4 IC<br />

Nd<br />

HR-ICP-MS<br />

Na + *<br />

IC<br />

Ni*<br />

HR-ICP-MS<br />

Br - *<br />

IC<br />

P* ICP-OES<br />

Cl - *<br />

IC<br />

Pb*<br />

HR-ICP-MS<br />

F - *<br />

IC<br />

Pd<br />

HR-ICP-MS<br />

-<br />

NO 2 IC<br />

Pt<br />

HR-ICP-MS<br />

2­<br />

NO 3 IC<br />

Rb<br />

HR-ICP-MS<br />

3­<br />

PO 4 IC<br />

S* ICP-OES<br />

SO 2- 4 *<br />

IC<br />

* Chemicals detected in flowback and produced water. See Table A-3 in Appendix A.<br />

Although the majority of the species that are being quantified in this study have been identified in<br />

flowback or produced water, 60 the species relationships and relative quantities of the species in<br />

60 See Table A-3 in Appendix A.<br />

97


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

other sources (i.e., coal-fired utility boiler and acid mine drainage discharges) will differ (Chapman<br />

et al., 2012). This will allow the models described below to distinguish among the contributions<br />

from each source type.<br />

5.1.3.2. Source Apportionment Modeling<br />

The EPA is using the data gathered through the analyses described above to support source<br />

apportionment modeling. This source apportionment effort will use peer-reviewed receptor models<br />

to identify and quantify the relative contribution of different contaminant source types to<br />

environmental samples. 61 In this case, river samples collected near PWS intakes are being evaluated<br />

to discern the contributing sources (e.g., hydraulic fracturing wastewater or acid mine drainage) of<br />

bromide and chloride to those stream waters. Receptor models require a comprehensive analysis of<br />

environmental samples to provide a sufficient number of constituents to identify and separate the<br />

impacts of different source types. Analysis of major ions and inorganic trace elements (Table 39)<br />

will accomplish the needs for robust receptor modeling. Contaminant sources may be distinguished<br />

by unique ranges of chemical species and their concentrations, and the models provide quantitative<br />

estimates of the source type contributions along with robust uncertainty estimates.<br />

EPA-implemented models and commercial off-the-shelf software are being used to analyze the data<br />

from this particular study (e.g., Unmix, Positive Matrix Factorization, chemical mass balance). These<br />

models have previously been used to evaluate a wide range of environmental data for air, soil, and<br />

sediments (Cao et al., 2011; Pancras et al., 2011; Soonthornnonda and Christensen, 2008), and are<br />

now being used for emerging issues, such as potential impacts to drinking water from hydraulic<br />

fracturing.<br />

5.1.4. Status and Preliminary Data<br />

The EPA completed the two-week spring, summer, and fall intensive sampling periods beginning on<br />

May 16, July 20, and September 19, 2012, respectively. The EPA collected 206, 198, and 209<br />

samples during the spring, summer, and fall intensives, consisting of WWTF-treated discharge,<br />

river samples, raw hydraulic fracturing wastewater, and acid mine drainage. The data quality<br />

objectives (US EPA, 2012q) of 80% valid sample collection were met for both the spring (>85%)<br />

and summer (>96%) measurement intensives. Preparation work for the extraction and filtration<br />

of spring intensive samples for inductively coupled plasma optical emission spectroscopy and<br />

high-resolution magnetic sector field inductively coupled plasma mass spectrometry is ongoing.<br />

Table 40 shows the median discharge concentrations of chloride, bromide, sulfate, sodium, and<br />

conductivity in effluent from the two monitored WWTFs (prior to discharge and dilution in the<br />

rivers) during the spring sampling period; Table 40 also shows the conductivity of the effluent.<br />

Median chloride and sodium concentrations at Discharge A (Allegheny River) were almost 50% less<br />

than concentrations found at Discharge B (Blacklick Creek). High levels of sodium chloride<br />

(>20,000 milligrams per liter) are present in the discharge from both facilities (A and B). Bromide<br />

concentrations are roughly 35% lower at Discharge A than Discharge B.<br />

61 The receptor model, Positive Matrix Factorization, was peer-reviewed in 2007 (version 1.1) and 2011 (version 4.2), and<br />

Unmix (version 5.0) underwent peer review in 2007.<br />

98


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table 40. Median concentrations of selected chemicals and conductivity of effluent treated and discharged from two<br />

wastewater treatment facilities that accept oil and gas wastewater. Discharge A is located on the Allegheny River and<br />

Discharge B is located on Blacklick Creek, both in Pennsylvania. The EPA collected samples beginning on May 16,<br />

2012.<br />

Measurement<br />

Median Concentration<br />

(milligrams per liter)<br />

Discharge A<br />

Discharge B<br />

Chloride 49,875 97,963<br />

Bromide 506 779<br />

Sulfate 679 976<br />

Sodium 20,756 38,394<br />

Conductivity (millisiemens per centimeter) 110 168<br />

The differences in the discharge concentrations are due to a combination of the treatment<br />

processes<br />

and<br />

unique regional chemical characteristics<br />

of oil and<br />

gas wastewater<br />

being treated at<br />

each<br />

of the facilities. Additionally, the<br />

discharge<br />

from the<br />

WWTFs<br />

is diluted<br />

into<br />

surface<br />

waters<br />

with very different median flows, with<br />

the<br />

USGS provisional<br />

median<br />

flows<br />

for the<br />

river<br />

sampling<br />

events<br />

reported as 15,158<br />

and<br />

2,531<br />

cubic feet per<br />

second<br />

for the<br />

Allegheny River<br />

in spring<br />

( May<br />

16–30, 2012)<br />

and summer<br />

(July 20–<br />

August<br />

3, 2012)<br />

, respectively<br />

( USGS, 2012a);<br />

and<br />

642 and<br />

35<br />

cubic feet per second for Blacklick Creek<br />

in<br />

spring ( May<br />

17–<br />

31, 2012)<br />

and summer<br />

( July<br />

21–August<br />

4, 2012), respectively<br />

(USGS, 2012b).<br />

The<br />

relative<br />

impact of these seasonal dilution scenarios from<br />

the WWTF<br />

discharges will<br />

be<br />

determined<br />

with<br />

the<br />

measured chemical species.<br />

5. 1. 5.<br />

Next<br />

Steps<br />

Analysis<br />

of<br />

field<br />

and<br />

source samples<br />

will continue in order to obtain the necessary data for source<br />

apportionment<br />

modeling.<br />

Once sample analyses<br />

are completed, data will be used<br />

as input<br />

to the<br />

receptor models described above to identify and<br />

quantify<br />

the<br />

sources of<br />

chloride and<br />

bromide<br />

at<br />

PWS intakes.<br />

5. 1. 6.<br />

Quality<br />

Assurance<br />

Summary<br />

The<br />

“ QAPP<br />

for<br />

Hydraulic Fracturing Wastewater Source Apportionment” was<br />

approved<br />

on April 17,<br />

2012<br />

(US<br />

EPA, 2012q).<br />

A TSA<br />

of the<br />

field sampling<br />

was<br />

conducted<br />

on<br />

May<br />

3, 2012, by<br />

the<br />

designated EPA QA Manager. There were two<br />

findings and<br />

two<br />

observations.<br />

The agreed-upon<br />

corrective actions<br />

were reported<br />

in writing<br />

to<br />

the researchers and management on May<br />

17, 2012,<br />

and<br />

have been<br />

implemented<br />

by<br />

the research<br />

team.<br />

One<br />

finding<br />

identified the<br />

need<br />

to<br />

verbally<br />

“ call back” measurement<br />

numbers<br />

between<br />

the sampler<br />

and scribe to<br />

confirm values when<br />

collecting<br />

short-term<br />

river<br />

measurements.<br />

The researchers<br />

instituted the verbal confirmation immediately in the field<br />

as suggested<br />

by the auditor. The<br />

second<br />

finding highlighted the need to accurately track<br />

the<br />

sample cooler<br />

temperature. A corrective action<br />

was implemented to improve the monitoring/recording of sample shipping cooler temperatures by<br />

ordering new National Institute of Standards and<br />

Technology<br />

traceable logging temperature<br />

loggers and keeping the loggers with<br />

the samples<br />

throughout the day<br />

in order to record<br />

accurate<br />

data of the temperatures<br />

at which the samples are stored and shipped. The new<br />

loggers were<br />

received and used in the field on May 8, 2012.<br />

99


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

During the audit, it was observed that the custody seals may not have offered a level of security<br />

necessary for the project. The field team had already identified this potential problem and had<br />

ordered different tamper-resistant seals before the field trip. The new seals (NIK Public Safety<br />

Tamperguard brand evidence tape) have been in use since they were received on May 10, 2012.<br />

The second observation during the audit was the need to document the reasoning of changes<br />

performed to standard operating procedures. The researchers have documented all the changes<br />

performed as well as the logic and reasoning of the changes in the field laboratory notebooks. Most<br />

modifications to the procedures were related to procedural adjustments made as a result of the<br />

field site characteristics, which were slightly different from the field site characteristics used to<br />

field-test the procedures in North Carolina. The documents also included updates to points of<br />

contact, references, and added text for clarification (e.g., river velocity measurements). Revisions<br />

reflecting these changes have been made to the QAPP and four SOPs based on the spring intensive<br />

field experience and the TSA. The revised version of the QAPP and four SOPs were approved on<br />

June 29, 2012. These updates do not impact the original data quality objectives.<br />

The researchers are following the QA procedures described in the QAPP and the standard operating<br />

procedures. In accordance to the QAPP, a TSA was performed on July 16 and 17, 2012, to evaluate<br />

laboratory operations. The designated EPA QA Manager reviewed the ion chromatography and<br />

high-resolution magnetic sector field inductively coupled plasma mass spectrometer analyses, data<br />

processing, storage, sample receiving and chain of custody procedures. The audit identified two<br />

observations and one best practice. One of the observations highlighted the need for a process that<br />

would ensure proper transcription of the data from the ion chromatography instrument to the<br />

report file. To reduce uncertainty and potential transcription errors, the analyst developed a<br />

process to export the data produced by the instrument in a text file instead of copying and pasting<br />

the data to a separate file. Another observation was the need to include performance evaluation<br />

samples in the analytical set. The performance evaluation samples will be analyzed in addition to<br />

the other quality controls already in place, which include blanks, duplicates, standard reference<br />

materials, and continuing calibration verification. The performance evaluation audit is being<br />

scheduled as specified in the QAPP. The blind performance evaluation samples will be analyzed<br />

with the regular samples and the data reported back to the QA Manager of the organization<br />

providing the blind performance evaluation samples. The best practice identified by the auditor<br />

was the tracking system, which uses a scanner and bar codes to track sampling bottles through the<br />

whole process: preparation, deployment to/from the field, sample analysis, and data reporting. The<br />

quality control (QC) procedures described in the QAPP have been followed in all instances. Besides<br />

the two TSAs performed and the performance evaluation audit, an ADQ is being coordinated by the<br />

designated EPA QA Manager. The source apportionment modeling will be described in a separate<br />

modeling QAPP. A TSA will be scheduled in 2013 for the modeling component of the study.<br />

100


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

5.2. Wastewater Treatability Studies<br />

5.2.1. Relationship to the Study<br />

The EPA is conducting laboratory experiments to assess the efficacy of conventional wastewater<br />

treatment processes on selected chemicals found in hydraulic fracturing wastewater to provide<br />

data to inform the research question posed in Table 41. The results of the water treatability<br />

experiments also complement the surface water modeling research project (see Section 4.2).<br />

Table 41. Secondary research questions addressed by the wastewater treatability laboratory studies.<br />

Water Cycle Stage<br />

Wastewater treatment and<br />

waste disposal<br />

Applicable Research Questions<br />

How effective are conventional POTWs and commercial treatment<br />

systems in removing organic and inorganic contaminants of concern<br />

in hydraulic fracturing wastewater<br />

5.2.2. Introduction<br />

Hydraulic fracturing wastewater, including flowback and produced water, is generally disposed of<br />

through underground injection in Class II UIC wells or treatment by a WWTF followed by surface<br />

water discharge. A generalized diagram for the onsite flow of water is given in Figure 24. A US<br />

Department of Energy report provides a state-by-state description of costs, regulations, and<br />

treatment/disposal practices for hydraulic fracturing wastes, including wastewater (Puder and Veil,<br />

2006).<br />

Wastewater may be treated at a WWTF, such as a POTW or centralized waste treatment facility<br />

(CWT). This project focuses on the efficacy of treatment processes at POTWs and CWTs, since<br />

discharge of treated wastewater to surface waters provides an opportunity for chemicals found in<br />

the effluent to be transported to downstream PWS intakes. This project will also explore treatment<br />

processes used for reuse of hydraulic fracturing wastewater.<br />

101


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Figure 24. Hydraulic fracturing wastewater flow in unconventional oil and gas extraction. Flowback and produced<br />

water (collectively referred to as “hydraulic fracturing wastewater”) is typically stored onsite prior to disposal or<br />

treatment. Hydraulic fracturing wastewater may be disposed of through Class II underground injection control (UIC)<br />

wells or through surface water discharge following treatment at wastewater treatment facilities, such as publicly<br />

owned treatment works or centralized waste treatment facilities. Wastewater may be treated on- or offsite prior to<br />

reuse in hydraulic fracturing fluids.<br />

5.2.2.1. Publicly Owned Treatment Works Treatment Processes<br />

Conventional POTW treatment processes are categorized into four groups: primary, secondary,<br />

tertiary, and advanced treatment. A generalized flow diagram is presented in Figure 25.<br />

Primary treatment processes remove larger solids and wastewater constituents that either settle or<br />

float. These processes include screens, weirs, grit removal, and/or sedimentation and flotation (e.g.,<br />

primary clarification). Secondary treatment processes typically remove biodegradable organics by<br />

using microbial processes (e.g., “bioreactor” in Figure 25) in fixed media (e.g., trickling filters) or in<br />

the water column (e.g., aeration basins). There is typically another settling stage in the secondary<br />

treatment process where suspended solids generated in the aeration basin are removed through<br />

settling (“secondary clarifier” in Figure 25). In some systems, tertiary or advanced treatment (“filter<br />

and UV disinfection” in Figure 25) may be applied as a polishing step to achieve a particular end use<br />

water quality (e.g., for reuse in irrigation).The POTW then discharges the treated effluent to surface<br />

water, if recycling or reuse is not intended. Solid residuals formed as byproducts of the treatment<br />

processes may contain metals, organics, and radionuclides that were removed from the water.<br />

Residuals are typically de-watered and disposed of via landfill, land application, or incineration.<br />

102


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Figure 25. Generalized flow diagram for conventional publicly owned works treatment processes. See the text for<br />

descriptions of primary, secondary, tertiary, and advanced treatment processes.<br />

The exact number of POTWs currently accepting hydraulic fracturing wastewater is not known. In<br />

Pennsylvania, where gas production from the Marcellus Shale is occurring, approximately 15<br />

POTWs were accepting hydraulic fracturing wastewater until approximately May 2011. In April<br />

2011, the Pennsylvania Department of Environmental Protection announced a request for<br />

Marcellus Shale natural gas drillers to voluntarily cease delivering their wastewater to the 15<br />

POTWs. The state also promulgated regulations in November 2011 that established monthly<br />

average limits (500 milligrams per liter TDS, 250 milligrams per liter chloride, 10 milligrams per<br />

liter total barium, and 10 milligrams per liter total strontium) for new and expanded TDS<br />

discharges (PADEP, 2011). These limits do not apply to the 15 facilities identified in the voluntary<br />

request or other grandfathered treatment plants.<br />

5.2.2.2. Commercial Waste Treatment Facility Processes<br />

Commercial processes for treating hydraulic fracturing wastewater include crystallization (zeroliquid<br />

discharge), thermal distillation/evaporation, electrodialysis, reverse osmosis, ion exchange,<br />

and coagulation/flocculation followed by settling and/or filtration. Some treatment processes are<br />

better able to treat high-TDS waters, which is a common property of hydraulic fracturing<br />

wastewater. Thermal processes are energy-intensive, but are effective at treating high-TDS waters<br />

103


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

and may be able to treat hydraulic fracturing wastewater with zero liquid discharge, leaving only a<br />

residual salt. Electrodialysis and reverse osmosis may be feasible for treating lower-TDS<br />

wastewaters. These technologies are not able to treat high-TDS waters (>45,000 milligrams per<br />

liter) and may require pre-treatment (e.g., coagulation and filtration) to minimize membrane<br />

fouling.<br />

Centralized waste treatment facilities can be used for pre-treatment prior to a POTW or, under an<br />

approved NPDES permit, can discharge directly to surface water (Figure 24). Commercial waste<br />

treatment processes will also result in some residual material that will require management and<br />

disposal.<br />

5.2.2.3. Reuse<br />

Gas producers are accelerating efforts to reuse and recycle hydraulic fracturing wastewater in some<br />

regions in order to decrease costs associated with procuring fresh water supplies, wastewater<br />

transportation, and offsite treatment and disposal. The EPA requested information on current<br />

wastewater management practices in the Marcellus Shale region from six oil and gas operators in<br />

May 2011. 62 Responses to the request for information indicated that reuse treatment technologies<br />

are similar, if not the same, to those used by WWTFs. Reuse technologies included direct reuse,<br />

onsite treatment (e.g., bag filtration, weir/settling tanks, third-party mobile treatment systems) and<br />

offsite treatment. Offsite treatment, in most instances, consisted of some form of stabilization,<br />

primary clarification, precipitation process, and secondary clarification and/or filtration. Specific<br />

details for offsite treatment methods were lacking as they are considered proprietary.<br />

Innovation in coupling various treatment processes may help reduce wastewater volumes and<br />

fresh water consumed in hydraulic fracturing operations. A challenge facing reuse technology<br />

development is treating water onsite to an acceptable quality for reuse in subsequent hydraulic<br />

fracturing operations. Key water quality parameters to control include TDS, calcium, and hardness,<br />

all of which play a major role in scale formation in wells.<br />

Recycling and reuse reduce the immediate need for treatment and disposal and water acquisition<br />

needs. There will likely be a need to treat and properly dispose of the final concentrated volumes of<br />

wastewater and residuals produced from treatment processes from a given area of operation,<br />

however.<br />

5.2.3. Research Approach<br />

The EPA is examining the fate and transport of chemicals through conventional POTW treatment<br />

processes and commercial chemical coagulation/settling processes. The objective of this work is to<br />

identify the partitioning of selected chemicals between solid and aqueous phases and to assess the<br />

biodegradation of organic constituents. In addition, microbial community health will be monitored<br />

in the reactors to identify the point where biological processes begin to fail. Contaminants that can<br />

pass through treatment processes and impact downstream PWS intakes will be identified.<br />

62 Documents received pursuant to the request for information are available at http://www.epa.gov/region3/<br />

marcellus_shale/.<br />

104


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Fate and Transport of Selected Contaminants in Wastewater Treatment Processes. The EPA will<br />

initially analyze the fate and transport of selected hydraulic fracturing–related contaminants in<br />

wastewater treatment processes, including conventional processes (primary clarifier, aeration<br />

basin, secondary clarifier), commercial processes (chemical precipitation/filtration and<br />

evaporation/distillation), and water reuse processes (pretreatment and filtration). The initial phase<br />

of this work will involve bench-scale fate and transport studies in a primary clarifier followed by 10<br />

liter chemostat reactors seeded with microbial organisms from POTW aeration basins. In benchscale<br />

work relevant to CWTs, similar fate and transport studies will be performed in chemical<br />

coagulation, settling, and filtration processes.<br />

A list of contaminants (Table 42) for initial treatability studies have been identified and are based<br />

on the list of hydraulic fracturing-related chemicals identified for initial analytical method<br />

development (Table 45 in Section 5.4). Table 42 may change as future information on toxicity and<br />

occurrence is gathered. In addition to monitoring the fate of the contaminants listed in Table 42 in<br />

treatment settings, impacts on conventional wastewater treatment efficiency will be monitored by<br />

examining changes in chemical oxygen demand, biological oxygen demand, and levels of nitrate,<br />

ammonia, phosphorus, oxygen, TDS, and total organic carbon in the aeration basin.<br />

105


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table 42. Chemicals identified for initial studies on the adequacy of treatment of hydraulic fracturing wastewaters by<br />

conventional publicly owned treatment works, commercial treatment systems, and water reuse systems. Chemicals<br />

were identified from the list of chemicals needing analytical method development (Table 45).<br />

Target Chemical<br />

CASRN<br />

2,2-Dibromo-3­<br />

nitrilopropionamide<br />

10222-01-2<br />

Acrylamide 79-06-1<br />

Arsenic* 7440-38-2<br />

Barium* 7440-39-3<br />

Benzene* 71-43-2<br />

Benzyl chloride 100-44-7<br />

Boron* 7440-42-8<br />

Bromide* 24959-67-9<br />

t-Butyl alcohol 75-65-0<br />

Chromium* 7440-47-3<br />

Diethanolamine 111-42-2<br />

Ethoxylated alcohols, C10–C14 66455-15-0<br />

Ethylbenzene* 100-41-4<br />

Ethylene glycol* 107-21-1<br />

Formaldehyde 82115-62-6<br />

Glutaraldehyde 111-30-8<br />

Target Chemical<br />

CASRN<br />

Isopropanol* 67-63-0<br />

Magnesium* 7439-95-4<br />

Manganese* 7439-96-5<br />

Methanol* 67-56-1<br />

Napthalene* 91-20-3<br />

Nonylphenol 68152-92-1<br />

Nonylphenol ethoxylate 68412-54-4<br />

Octylphenol 1806-26-4<br />

Octylphenol ethoxylate 26636-32-8<br />

Potassium* 7440-09-7<br />

Radium* 7440-14-4<br />

Sodium* 7440-23-5<br />

Strontium* 7440-24-6<br />

Thiourea 62-56-6<br />

Toluene* 108-88-3<br />

Uranium 7440-61-1<br />

Iron* 7439-89-6 Xylene* 1330-20-7<br />

* Chemicals reported to be in flowback and produced water. See Table A-3 in Appendix A.<br />

Characterization of Contaminants in Hydraulic Fracturing Wastewater Treatment Residuals. The EPA<br />

will examine the concentrations and chemical speciation of inorganic contaminants in treatment<br />

residuals. Residuals generated from the research described above will be analyzed for inorganic<br />

contaminant concentrations via EPA Method 3051A (Microwave Assisted Digestion) and<br />

inductively coupled argon plasma-optical emission spectrometry. Samples will also undergo<br />

analysis via X-ray absorption spectroscopy in order to assess oxidation state and chemical<br />

speciation of target contaminants. Organic contaminants will be analyzed via liquid or gas<br />

chromatography-mass spectrometry after accelerated solvent extraction of the solids.<br />

5.2.4. Status and Preliminary Data<br />

This research is currently in the planning stage.<br />

5.2.5. Next Steps<br />

Initial studies will focus on establishing thresholds of TDS tolerance in chemostat bioreactors. Once<br />

the basic salt thresholds have been established, selected chemicals from the 26R forms will be<br />

added to the salt stock solutions. Salt concentrations will be kept below the thresholds where<br />

effects on the biological processes were observed. Potentially biodegradable pollutants (e.g.,<br />

organics) will be measured, and the EPA will attempt to identify breakdown products.<br />

Constituents that are not biodegradable (e.g., elements and anions) will be tracked through the<br />

treatment process by analyzing system effluent using the appropriate EPA Methods and by<br />

106


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

analyzing residuals from the primary clarifier and the bioreactors. The results of these bench-scale<br />

studies will be applied to a pilot-scale system that would target compounds identified in benchscale<br />

studies as being the most problematic due to their lack of degradation or removal in the<br />

treatment process.<br />

For studies on commercial treatment systems using chemical addition/settling, the EPA plans to<br />

conduct jar tests that employ coagulants/flocculants at appropriate contact and settling times. The<br />

jar tests will be conducted at the bench-scale using actual hydraulic fracturing wastewater samples.<br />

The EPA will also attempt to mimic evaporative/distillation processes by using thermal treatment<br />

on actual hydraulic fracturing wastewater samples. Both the jar test samples and residuals from<br />

thermal treatment will be analyzed for the chemicals listed in Table 42. Elements in the residuals<br />

will also be characterized via X-ray diffraction and X-ray absorption microscopy.<br />

5.2.6. Quality Assurance Summary<br />

The initial QAPP, “Fate, Transport and Characterization of Contaminants in Hydraulic Fracturing<br />

Water in Wastewater Treatment Processes,” was submitted on December 20, 2011, and approved<br />

in August 2012 (US EPA, 2012q).<br />

Because project activities are still in an early stage, no TSA has been performed. A TSA will be<br />

performed once the project advances to the data collection stage.<br />

As results are reported and raw data are provided from the laboratories, ADQs will be performed to<br />

verify that the quality requirements specified in the approved QAPP were met. Data will be<br />

qualified if necessary, based on these ADQs. The results of the ADQs will be reported with the<br />

summary of results in the final report.<br />

5.3. Brominated Disinfection Byproduct Precursor Studies<br />

The EPA is assessing the ability of hydraulic fracturing wastewater to contribute to DBP formation<br />

in drinking water treatment facilities, with a particular focus on the formation of brominated DBPs.<br />

This work will inform the following research question listed in Table 43 and is complemented by<br />

the analytical method development for DBPs (see Section 5.4).<br />

Table 43. Secondary research questions potentially answered by studying brominated DBP formation from treated<br />

hydraulic fracturing wastewater.<br />

Water Cycle Stage<br />

Wastewater treatment and<br />

waste disposal<br />

Applicable Research Questions<br />

What are the potential impacts from surface water disposal of treated<br />

hydraulic fracturing wastewater on drinking water treatment<br />

facilities<br />

5.3.1. Introduction<br />

Wastewaters from hydraulic fracturing processes typically contain high concentrations of TDS,<br />

including significant concentrations of chloride and bromide. These halogens are difficult to remove<br />

from wastewater; if discharged from treatment works, they can elevate chloride and bromide<br />

concentrations in drinking water sources. Upon chlorination at a drinking water treatment facility,<br />

chloride and bromide can react with naturally occurring organic matter (NOM) in the water and<br />

107


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

lead to the formation of DBPs. Because of their carcinogenicity and reproductive and<br />

developmental affects, the maximum contaminant levels (MCLs) of the DBPs bromate, chlorite,<br />

haloacetic acids, and total THMs in finished drinking water are regulated by the National Primary<br />

Drinking Water Regulations. 63 Table 44 summarizes the DBPs regulated and their corresponding<br />

MCLs.<br />

Increased bromide concentrations in drinking water resources can lead to greater total THM<br />

concentrations on a mass basis and may make it difficult for some PWSs to meet the regulatory<br />

limits of total THM listing in Table 44 in finished drinking water. As a first step, this project is<br />

examining the formation of brominated THMs, including bromoform (CHBr 3),<br />

dibromochloromethane (CHClBr 2), and bromodichloromethane (CHCl 2Br), during drinking water<br />

treatment processes. The formation of haloacetic acids (HAAs) and nitrosamines during drinking<br />

water treatment processes is also being investigated. 64<br />

Reactions of brominated biocides used in hydraulic fracturing operations with typical drinking<br />

water disinfectants associated with chlorination or chloramination are also being explored. 65<br />

Brominated biocides are often used in fracturing fluids to minimize biofilm growth. The objective of<br />

this work is to assess the contribution, if any, to brominated DBP formation and identify<br />

degradation pathways for brominated biocides.<br />

63 Authorized by the Safe Drinking Water Act.<br />

64 Nitrosamines are byproducts of drinking water disinfection, typically chloramination, and currently unregulated by the<br />

EPA. Data collected from the second Unregulated Contaminant Monitoring Rule indicate that nitrosamines are frequently<br />

being found in PWSs. Nitrosamines are potentially carcinogenic.<br />

65 Chlorination and chloramination are common disinfection processes used for drinking water.<br />

108


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table 44. Disinfection byproducts regulated by the National Primary Drinking Water Regulations.<br />

Disinfection Byproduct<br />

Maximum Contaminant Level Maximum Contaminant Level †<br />

Goal* (milligrams per liter)<br />

(milligrams per liter)<br />

Total Trihalomethanes<br />

Bromodichloromethane<br />

Zero<br />

Bromoform<br />

Zero<br />

0.080 as an annual average<br />

(sum of the concentrations of all<br />

Dibromochloromethane<br />

0.060<br />

four trihalomethanes)<br />

Chloroform 0.070<br />

Haloacetic Acids<br />

Dichloroacetic acid<br />

Zero<br />

Trichloroacetic acid 0.020<br />

Monochloroacetic acid 0.070 0.060 as an annual average<br />

Bromoacetic acid<br />

Regulated with this group but has (sum of the concentrations of all<br />

no MCL goal<br />

five haloacetic acids)<br />

Dibromoacetic acid<br />

Regulated with this group but has<br />

no MCL goal<br />

Bromate Zero 0.010 as an annual average<br />

Chlorite 0.80 1.0<br />

* A maximum contaminant level goal is the non-enforceable concentration of a contaminant in drinking water below<br />

which there is no known or expected risk to health; they are established under the Safe Drinking Water Act.<br />

†<br />

A maximum contaminant level (MCL) is an enforceable standard corresponding to the highest level of a<br />

contaminant that is allowed in drinking water. MCLs are set as close to MCL goals as feasible using the best<br />

available treatment technology and taking cost into consideration. MCLs are set under the Safe Drinking Water Act<br />

and apply only to water delivered by public water supplies (water supplies that serve 15 or more service connections<br />

or regularly serves an average of 25 or more people daily at least 60 days out of the year) (40 CFR 141.2).<br />

It is important to note that hydraulic fracturing wastewater can potentially contain other<br />

contaminants in significant concentrations that could affect human health. The EPA identified the<br />

impacts of elevated bromide and chloride levels in surface water from hydraulic fracturing<br />

wastewater discharge as a priority for protection of public water supplies. This project will<br />

ultimately provide PWSs with information on the potential for brominated DBP formation in<br />

surface waters receiving discharges from WWTFs.<br />

5.3.2. Research Approach<br />

This research will (1) analyze and characterize hydraulic fracturing wastewater for presence of<br />

halides, (2) evaluate the effects of high TDS upon chlorination of surface water receiving discharges<br />

of treated hydraulic fracturing wastewater, and (3) examine the reactions of brominated biocides<br />

subjected to chlorination during drinking water treatment. Selected analytes for characterizing<br />

hydraulic fracturing wastewater include nitrosamines and the halide anions chloride, bromide, and<br />

iodide—ions that are the likeliest to form DBPs (Richardson, 2003), including THMs and HAAs.<br />

Hydraulic fracturing wastewater samples have been obtained from several sources in Pennsylvania.<br />

The quantification of background concentrations of halides in the samples follows EPA Method<br />

300.1 (rev. 1) and the modified version of the method using mass spectrometry detection for<br />

bromide and bromate (discussed in Section 5.4). The samples are also being analyzed for the<br />

presence of DBPs, including THMs (EPA Method 551.1), HAAs (EPA Method 552.1), and N­<br />

109


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

nitrosamines (EPA Method 521), as well as elemental composition, anion concentration, TDS, and<br />

total organic carbon.<br />

Three treatments are being applied to high-TDS wastewater samples: (1) samples will be blended<br />

with deionized water at rates that mimic discharge into varying flow rates of receiving water in<br />

order to account for dilution effects; (2) samples will be blended with deionized water with NOM<br />

additions at concentration ranges typically found in surface waters; and (3) samples will be<br />

blended with actual surface water samples from rivers that receive treated hydraulic fracturing<br />

wastewater discharges. All samples will be subjected to formation potential experiments in the<br />

presence of typical drinking water disinfectants associated with chlorination or chloramination.<br />

Formation potential measures will be obtained separately for THMs, HAAs, and nitrosamines.<br />

Disinfection byproduct formation in surface water samples will be compared with DBP formation in<br />

deionized water as well as deionized water fortified with several NOM isolates from different water<br />

sources in order to examine the effects of different NOM on DBP formation. 66<br />

The brominated biocides 2,2-dibromo-3-nitropropionamide and 2-bromo-2-nitrol-1,3-propanediol,<br />

employed in hydraulic fracturing processes, are being subjected to chlorination conditions<br />

encountered during drinking water treatment. These experiments should provide insight on the<br />

potential formation of brominated THMs from brominated biocides. Effects of chlorination on the<br />

brominated biocides are also being monitored.<br />

5.3.3. Status and Preliminary Data<br />

Work has begun on total THM formation studies to identify potential problems with analysis (EPA<br />

Method 551.1) due to the high TDS levels typical in hydraulic fracturing wastewater. Wastewater<br />

influent and effluent samples were obtained from researchers involved in the source<br />

apportionment studies (Section 5.1) at two CWTs in Pennsylvania that are currently accepting<br />

hydraulic fracturing wastewater for treatment via chemical addition and settling. For this<br />

preliminary research, samples were diluted 1:100 with deionized water and equilibrated with<br />

sodium hypochlorite until a 2 milligrams per liter concentration of sodium hypochlorite was<br />

achieved (a typical disinfectant concentration for finished water from a PWS). The samples are<br />

being analyzed for pH, metals, TDS, total suspended solids, total organic content, and selected<br />

anions.<br />

Efforts to identify and quantify the parent brominated biocides using liquid chromatography/mass<br />

spectrometry methods have been unsuccessful to date, possibly due to poor ionization of the<br />

brominated molecules. The biocide samples subject to chlorination have been prepared for analysis<br />

of THMs.<br />

66 The concentration, chemical composition, and reactivity of NOM varies by geographic location due to factors such as<br />

presence and type of vegetation, physical and chemical properties of the surrounding soil and water, biological activity,<br />

and human activity among many others.<br />

110


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

5.3.4. Next Steps<br />

When the preliminary work on potential analytical effects from high TDS on total THM recovery is<br />

complete, a series of experiments to assess the potential formation of DBPs during chlorination will<br />

be run on the following samples:<br />

• Deionized water<br />

• Deionized water, varying concentrations of NOM<br />

• Deionized water plus TDS<br />

• Deionized water plus TDS and NOM<br />

• Hydraulic fracturing wastewater<br />

This series of samples will allow THM formation comparisons between hydraulic fracturing<br />

wastewater samples and less complex matrices. Dilutions will be made on the samples based on<br />

effluent discharge rates for existing WWTFs and receiving water flow rates. The samples will<br />

undergo chlorination and be sub-sampled over time (e.g., 0 to 120 minutes). Chloride to bromide<br />

ratios will be set at 50:1, 100:1, and 150:1 to encompass the range of conditions that may be found<br />

in surface waters impacted by varying concentrations of chloride and bromide. The sub-samples<br />

will be analyzed for individual THMs and formation kinetics will be determined. The EPA<br />

anticipates obtaining data for the formation of HAAs and nitrosamines, though THMs are the<br />

priority at this time.<br />

5.3.5. Quality Assurance Summary<br />

The initial QAPP, “Formation of Disinfection By-Products from Hydraulic Fracturing Fluids,” was<br />

submitted on June 28, 2011, and approved on October 5, 2011 (US EPA, 2011h). On June 7, 2012, an<br />

addendum was submitted and approved on June 28, 2012; this provided more details on<br />

modifications to EPA Method 300.1 for optimizing bromide/bromate recoveries in high-salt<br />

matrices. There are no deviations from existing QAPPs to report at this time.<br />

A TSA was performed on March 15, 2012, for this research project. Five findings were observed,<br />

related to improved communication, project documentation, sample storage, and QA/QC checks.<br />

Recommended corrective actions were accepted to address the findings. Since the TSA was<br />

performed before data generation activities, no impact on future reported results is expected. It is<br />

anticipated that a second TSA will be performed as the project progresses.<br />

As raw data are provided from the laboratories and results are reported, ADQs will be performed to<br />

verify that the quality requirements specified in the approved QAPP have been met. Data will be<br />

qualified if necessary based on these ADQs. Audits of data quality are scheduled for the first quarter<br />

of 2013 (none have been performed yet). The results of these ADQs will be reported with the<br />

summary of results in the final report.<br />

111


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

5.4. Analytical Method Development<br />

5.4.1. Relationship to the Study<br />

Sample analysis is an integral part of the EPA’s Plan to Study the Potential Impacts of Hydraulic<br />

Fracturing on Drinking Water Resources (US EPA, 2011e) and is clearly specified in research plans<br />

being carried out for the study’s retrospective case studies, prospective case studies, and laboratory<br />

studies. The EPA requires robust analytical methods to accurately and precisely determine the<br />

composition of hydraulic fracturing-related chemicals in ground and surface water, flowback and<br />

produced water, and treated wastewater.<br />

5.4.2. Project Introduction<br />

Analytical methods enable accurate and precise measurement of the presence and quantities of<br />

different chemicals in various matrices. Since the quantification of the presence or absence of<br />

hydraulic fracturing-related chemicals will likely have substantial implications for the conclusions<br />

of the study, it is important that robust analytical methods exist for chemicals of interest.<br />

In many cases, standard EPA methods that have been designed for a specific matrix or set of<br />

matrices can be used for this study. Standard EPA methods are peer-reviewed and officially<br />

promulgated methods that are used under different EPA regulatory programs. For example, EPA<br />

Method 551.1 is being used to detect THMs as part of the Br-DBP research project (see Section 5.3)<br />

and EPA Method 8015D is being used to detect diesel range organics in ground and surface water<br />

samples collected as part of the retrospective case studies (see Chapter 7).<br />

In other cases, standard EPA methods are nonexistent for a chemical of interest. In these situations,<br />

methods published in the peer-reviewed literature or developed by consensus standard<br />

organizations (e.g., the American Society for Testing and Materials, or ASTM) are used. However,<br />

these methods are rarely developed for or tested within matrices associated with the hydraulic<br />

fracturing process. In rare, but existing cases, where no documented methods exist, researchers<br />

generally develop their own methods for determining the concentrations of certain chemicals of<br />

interest. For these latter two situations, the analytical methods chosen must undergo rigorous<br />

testing, verification, and potential validation to ensure that the data generated they generate are of<br />

known and high quality. The EPA has identified selected chemicals found in hydraulic fracturing<br />

fluids and wastewater for the development and verification of analytical methods.<br />

5.4.3. Research Approach<br />

5.4.3.1. Chemical Selection<br />

Hydraulic fracturing-related chemicals include chemicals used in the injected fracturing fluid,<br />

chemicals found in flowback and produced water, and chemicals resulting from the treatment of<br />

hydraulic fracturing wastewater (e.g., chlorination or bromination at wastewater treatment<br />

facilities). Some of these chemicals are present due to the mobilization of naturally occurring<br />

chemicals within the geologic formations or through the degradation or reaction of the injected<br />

chemicals in the different environments (i.e., subsurface, surface and wastewater). The EPA has<br />

identified over 1,000 chemicals that are reported to be used in fracturing fluids or found in<br />

hydraulic fracturing wastewaters (see Appendix A); these range from the inert and innocuous, such<br />

as sand and water, to reactive and toxic chemicals, like alkylphenols and radionuclides.<br />

112


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

To help choose chemicals for analytical method testing, a group of EPA researchers and analytical<br />

laboratory chemists discussed the factors most important to their research needs and to the overall<br />

study. The following criteria were developed to identify a subset of the chemicals listed in Appendix<br />

A for initial analytical method testing activities:<br />

• Frequency of occurrence 67 in hydraulic fracturing fluids and wastewater<br />

• Toxicity 68<br />

• Mobility in the environment (expected fate and transport)<br />

• Availability of instrumentation/detection systems for the chemical<br />

Table 45 lists the chemicals selected for analytical method testing and development. It includes 14<br />

different classes of chemicals, 51 specifically identified elements or compounds, six groups of<br />

compounds (e.g., ethoxylated alcohols and light petroleum distillates), and two related physical<br />

properties (gross α and gross β analyses associated with radionuclides). The EPA will continually<br />

review Table 45 and add new chemicals as needed.<br />

67 Occurrence information was gathered from the US House of Representatives report Chemicals Used in Hydraulic<br />

Fracturing (2011) (USHR, 2011)and Colborn et al. (2011). Chemicals with high frequencies were considered for inclusion.<br />

However, some high-frequency chemicals were ultimately not included in the EPA’s priority list of chemicals of interest.<br />

For example, while silica or silicon dioxide is often near the top of lists in terms of frequency of occurrence, this likely<br />

refers to the sand that is used as a proppant during the hydraulic fracturing process. Additionally, certain chemicals, such<br />

as hydrogen chloride or sulfuric acid, no longer exist as the initial compounds once dissolved in water and often react<br />

with other compounds. As a result, these chemicals, and others, were not added to the list.<br />

68 Colborn et al. (2011) provided toxicity information compiled from MSDS from industry and government agencies and<br />

compared the chemicals in their list with toxic chemical databases, such as TOXNET and the Hazardous Substances<br />

Database.<br />

113


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table 45. Chemicals identified for analytical method testing activities. Selection criteria for the chemicals included, but were not limited to, frequency of occurrence<br />

in fracturing fluids and wastewater, toxicity, environmental mobility, and availability of detection systems for the chemical.<br />

Chemical Class Chemical Name(s) CASRN Purpose in Hydraulic Fracturing Reason Selected<br />

Propargyl alcohol 107‐19‐7<br />

Alcohols<br />

Methanol 67‐56‐1 Corrosion inhibitor<br />

Isopropanol 67‐63‐0<br />

Toxicity, frequency of use<br />

t‐Butyl alcohol 75‐65‐0 Byproduct of t‐butyl hydroperoxide<br />

Aldehydes<br />

Glutaraldehyde 111‐30‐8 Biocide<br />

Formaldehyde 50‐00‐0 Biocide<br />

Toxicity, frequency of use<br />

Alkylphenols<br />

Octylphenol 27193-28-8<br />

Nonylphenol 84852-15-3<br />

Surfactant<br />

Toxicity, frequency of use<br />

Alkylphenol Octylphenol ethoxylate 9036-19-5<br />

ethoxylates Nonylphenol ethoxylate 26027-38-3<br />

Surfactant<br />

Frequency of use<br />

Thiourea 62‐56‐6 Corrosion inhibitor Toxicity<br />

Amides<br />

Acrylamide 79‐06‐1 Friction reducer Toxicity, frequency of use,<br />

requested by EPA<br />

2,2‐Dibromo‐3‐nitrilopropionamide 10222‐01‐2 Biocide<br />

researchers<br />

Amines (alcohol) Diethanolamine 111-42-2 Foaming agent Frequency of use<br />

Aromatic<br />

hydrocarbons<br />

BTEX, naphthalene, benzyl<br />

chloride, light petroleum<br />

hydrocarbons<br />

Carbohydrates Polysaccharides Byproduct<br />

Disinfection<br />

byproducts<br />

Ethoxylated<br />

alcohols<br />

Trihalomethanes, haloacetic acids,<br />

N-nitrosamines*<br />

Ethoxylated alcohols,<br />

C8–10 and C12–18<br />

Gelling agents, solvents<br />

Byproduct<br />

Toxicity, frequency of use,<br />

requested by EPA<br />

researchers<br />

Requested by EPA<br />

researchers<br />

Toxicity<br />

68954-94-9 Surfactant Frequency of use<br />

Table continued on next page<br />

114


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

Chemical Class Chemical Name(s) CASRN Purpose in Hydraulic Fracturing Reason Selected<br />

Ethylene glycol 107‐21‐1<br />

Diethylene glycol 111‐46‐6<br />

Crosslinker, breaker, scale inhibitor<br />

Triethylene glycol 112-27-6<br />

Glycols<br />

Frequency of use<br />

Tetraethylene glycol 112-60-7<br />

2‐Methoxyethanol † 109‐86‐4<br />

2-Butoxyethanol † 111-76-2<br />

Foaming agent<br />

Halogens Chloride 16887‐00‐6 Brine carrier fluid, breaker Frequency of use<br />

Barium 7440‐39‐3 Mobilized during hydraulic fracturing<br />

Inorganics<br />

Strontium 7440‐24‐6 Mobilized during hydraulic fracturing Toxicity, frequency of use<br />

of potassium and sodium<br />

Boron 7440‐42‐8 Crosslinker<br />

salts, mobilization of<br />

Sodium 7440‐23‐5 Brine carrier fluid, breaker<br />

naturally occurring ions<br />

Potassium 7440‐09‐7 Brine carrier fluid<br />

Gross α<br />

Gross β<br />

Toxicity, mobilization of<br />

Radionuclides Radium 13982‐63‐3 Mobilized during hydraulic fracturing<br />

naturally occurring ions<br />

Uranium 7440‐61‐1<br />

Thorium 7440‐29‐1<br />

* See Section 5.3.<br />

†<br />

These compounds are chemically similar to glycols and are analyzed using the same methods.<br />

115


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

5.4.3.2. Analytical Method Testing and Development<br />

Method Development. The EPA’s process for analytical method development is shown in Figure 26.<br />

In the first step, an existing base method is identified for the specific chemical(s) of interest in a<br />

given matrix. Base methods may include promulgated, standard methods or, if no standard<br />

methods are available, methods existing in peer-reviewed literature or developed through a<br />

consensus standard organization.<br />

Select base<br />

analytical method<br />

Accept base method<br />

Yes<br />

QA/QC testing:<br />

Does the method meet<br />

specified QA/QC acceptance<br />

criteria<br />

No<br />

Identify potential<br />

reasons for QA/QC failing<br />

acceptance criteria<br />

Accept modified base<br />

method and prepare<br />

standard operating<br />

procedure (SOP)<br />

Modify method to correct<br />

cause of interference<br />

No<br />

Major modifications<br />

to method<br />

Yes<br />

Repeat testing:<br />

Does the modified method<br />

meet specified QA/QC<br />

acceptance criteria<br />

No<br />

Modify method to<br />

correct cause of<br />

failure to meet criteria<br />

Yes<br />

No, repeatedly<br />

Prepare SOP for<br />

method<br />

Develop new method if<br />

existing method cannot be<br />

modified to meet QA/QC<br />

acceptable criteria<br />

Figure 26. Flow diagram of the EPA’s process leading to the development of modified or new analytical methods.<br />

Analytical methods may exist for specific chemicals or for a general class of chemicals (e.g.,<br />

alcohols). Table 46 lists the base methods identified for the 14 chemical classes shown in Table 45.<br />

116


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table 46. Existing standard methods for analysis of selected hydraulic fracturing-related chemicals listed in Table 45.<br />

The EPA will analyze samples using existing methods to determine if the procedure meets the quality assurance<br />

criteria for the current study.<br />

Chemical Class<br />

Standard Method*<br />

Alcohols<br />

SW-846 Methods 5030 and 8260C<br />

Aldehydes SW-846 Method 8315<br />

Alkylphenols<br />

No standard method<br />

Alkylphenol ethoxylates<br />

No standard method<br />

Amides<br />

SW-846 Methods 8032A<br />

Amines (alcohols)<br />

No standard method<br />

Aromatic hydrocarbons<br />

SW-846 Methods 5030 and 8260C<br />

Carbohydrates<br />

No standard method<br />

Disinfection byproducts DWA Methods 521, 551, and 552<br />

Ethoxylated alcohols ASTM D7485-09<br />

Glycols<br />

Region 3 Draft Standard Operating Procedure<br />

Halogens<br />

SW-846 Method 9056A<br />

Inorganic elements<br />

SW-846 Methods 3015A and 6020A<br />

Radionuclides SW-846 Method 9310<br />

* DWA methods can be found at http://water.epa.gov/scitech/methods/cwa/index.cfm. SW-846<br />

Methods can be found at http://www.epa.gov/epawaste/hazard/testmethods/sw846/online/<br />

index.htm.<br />

Once a candidate base method is selected, 69 an initial QA/QC round of testing is conducted. Testing<br />

occurs first with spiked laboratory water samples to familiarize the analyst with the method<br />

procedure, eliminate any potential matrix interferences, and determine various QA/QC control<br />

parameters, such as sensitivity, bias, precision, spike recovery, and analytical carry-over potential<br />

(sample cross-contamination). The results from the initial QA/QC testing are examined to<br />

determine if they meet the acceptance criteria specified in the QAPP (US EPA, 2011g) and thus are<br />

sufficient to meet the needs of the research study. Some of the key QA/QC samples examined<br />

include:<br />

• Standard and certified reference materials (where available) for bias<br />

• Matrix and surrogate spikes for bias (when reference materials are not available) and<br />

matrix interferences<br />

• Replicates for precision<br />

• Blanks for analytical carry-over<br />

If an acceptance criterion for any of the QA/QC samples is not met, the sample is typically re-run to<br />

ensure that the result is not a random event. If an acceptance criterion is repeatedly not met, a<br />

69 Additional information on selecting a base method can be found in the QAPP, “Quality Assurance Project Plan for the<br />

Chemical Characterization of Select Constituents Relevant to Hydraulic Fracturing,” found at<br />

http://www.epa.gov/<strong>hf</strong>study/qapps.html.<br />

117


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

systematic problem is indicated, and method modification is undertaken to help reduce or<br />

eliminate the problem.<br />

The method modification process can take many forms, depending on the specific circumstances,<br />

and may include changing sample preparation and cleanup techniques, solvents, filters, gas flow<br />

rates, temperature regimes, injector volumes, chromatographic columns, analytical detectors, etc.<br />

Once the method modification process is complete, the analysis is repeated as described above<br />

using spiked laboratory water samples. If the new QA/QC sample results meet the acceptance<br />

criterion, the method modification is deemed to have been successful for that matrix and an<br />

updated SOP is prepared. Additional testing in more complex water matrices will continue, if<br />

appropriate.<br />

If testing and modification of the identified base method fails to accurately and precisely quantify<br />

the chemical of interest and/or fails to have the sensitivity required by the research program, the<br />

EPA may undertake new method development activities.<br />

Method Verification. Method verification determines the robustness of successfully tested and<br />

modified analytical methods. This involves the preparation of multiple blind spiked samples (i.e.,<br />

samples whose concentrations are only known to the sample preparer) by an independent chemist<br />

(i.e., one not associated with developing the method under testing and verification) and the<br />

submission of the samples to at least three other analytical laboratories participating in the<br />

verification process. Results from the method verification process can lead to either the acceptance<br />

of the method or re-evaluation and further testing of the method (US EPA, 1995).<br />

Method Validation. The final possible step in analytical method testing and development is method<br />

validation. Method validation involves large, multi-laboratory, round robin studies and is generally<br />

conducted by the EPA program offices responsible for the publication and promulgation of<br />

standard EPA methods.<br />

5.4.4. Status, Preliminary Data, and Next Steps<br />

Method development, testing, and verification are being conducted according to the procedures<br />

outlined in two QAPPs: “Quality Assurance Project Plan for the Chemical Characterization of Select<br />

Constituents Relevant to Hydraulic Fracturing” (US EPA, 2011g) and “Quality Assurance Project<br />

Plan for the Inter-Laboratory Verification and Validation of Diethylene Glycol, Triethylene Glycol,<br />

Tetraethylene Glycol, 2-Butoxyethanol and 2-Methoxyethanol in Ground and Surface Waters by<br />

Liquid Chromatography/Tandem Mass Spectrometry” (US EPA, 2012r).<br />

5.4.4.1. Glycols and Related Compounds<br />

Glycols (diethylene glycol, triethylene glycol, and tetraethylene glycol) and the chemically related<br />

compounds 2-butoxyethanol and 2-methoxyethanol are frequently used in hydraulic fracturing<br />

fluids and not naturally found in ground water. Thus, they may serve as reliable indicators of<br />

contamination of ground water from hydraulic fracturing activities. EPA Method 8015b is the gas<br />

chromatography-flame ionization detector method typically used to analyze for glycols; however,<br />

the sensitivity is not sufficient for the low-level analysis required for this project. Therefore, the<br />

EPA’s Region 3 Environmental Science Center developed a method for the determination and<br />

118


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

quantification of these compounds using liquid chromatography-tandem mass spectrometry. The<br />

method is based on ASTM D7731-11e1 and EPA SW-846 Method 8321. The EPA is currently<br />

verifying this method to determine its efficacy in identifying and quantifying these compounds in<br />

drinking water and other water matrices associated with the hydraulic fracturing process.<br />

5.4.4.2. Acrylamide<br />

Acrylamide is often used as a friction reducer in injected hydraulic fracturing fluids (GWPC,<br />

2012b). EPA SW-846 Methods 8316 and 8032A are both suitable methods for the analysis of<br />

acrylamide. Method 8316 involves analysis by high-performance liquid chromatography with<br />

ultraviolet detector at 195 nanometers, with a detection level of 10 micrograms per liter. This<br />

short wavelength, however, is not very selective for acrylamide (i.e., interferences are likely), and<br />

the sensitivity is not adequate for measurements in water. Method 8032A involves the<br />

bromination of acrylamide, followed by gas chromatography-mass spectrometry analysis. This<br />

method is much more selective for acrylamide, and detection limits are much lower (0.03<br />

micrograms per liter). However, in complex matrices (e.g., hydraulic fracturing wastewater), the<br />

accuracy and precision of acrylamide analysis may be limited by poor extraction efficiency and<br />

matrix interference.<br />

To avoid reactions with other compounds present in environmental matrices and to lower the<br />

detection limit, the EPA is developing a new analytical method for the determination of acrylamide<br />

at very low levels in water containing a variety of additives. The method currently under<br />

development involves solid phase extraction with activated carbon followed by quantitation by<br />

liquid chromatography-tandem mass spectrometry using an ion exclusion column. The EPA has<br />

begun the multi-laboratory verification of the method.<br />

5.4.4.3. Ethoxylated Alcohols<br />

Surfactants are often added to hydraulic fracturing fluids to decrease liquid surface tension and<br />

improve fluid passage through pipes. Most of the surfactants used are alcohols or some derivative<br />

of an ethoxylated compound, typically ethoxylated alcohols. Many ethoxylated alcohols and<br />

ethoxylated alkylphenols biodegrade in the environment, but often the degradation byproducts<br />

are toxic (e.g., nonylphenol, a degradation product of nonylphenol ethoxylate, is an endocrine<br />

disrupting compound) (Talmage, 1994). No standard method currently exists for the<br />

determination of ethoxylated alcohols; therefore, the EPA is developing a quantitative method for<br />

ethoxylated alcohols. ASTM Method D 7458-09 and USGS Method Number O1433-01 were used<br />

as starting points for this method development effort; both of these methods involve solid-phase<br />

extraction followed by liquid chromatography-tandem mass spectrometry quantitation. These<br />

methods both allow the analysis of nonylphenol diethoxylate and alkylphenols, but there are<br />

currently no standard methods for the analysis of the full range of nonylphenol ethoxylate<br />

oligomers (EO 3–EO 20) or alcohol ethoxylate oligomers (C 12–15EO x, where x = 2–20). This method<br />

SOP is being prepared and will be followed by method verification.<br />

5.4.4.4. Disinfection Byproducts<br />

Flowback and produced water can contain high levels of TDS, which may include bromide and<br />

chloride (US EPA, 2012d). In some cases, treatment of flowback and produced water occurs at<br />

WWTFs, which may be unable to effectively remove bromide and chloride from hydraulic<br />

119


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

fracturing wastewater before discharge. The presence of bromide ions in source waters undergoing<br />

chlorination disinfection may lead to the formation of brominated DBPs—including bromate, THMs,<br />

and HAAs—upon reaction with natural organic material (Richardson, 2003). Brominated DBPs are<br />

considerably more toxic than corresponding chlorinated DBPs (Plewa et al., 2004; Richardson et al.,<br />

2007) and have higher molecular weight. Therefore, on an equal molar basis, brominated DBPs will<br />

have a greater concentration by weight than chlorinated DBPs, hence leading to a greater likelihood<br />

of exceeding the total THM and HAA MCLs that are stipulated in weight concentrations (0.080 and<br />

0.060 milligrams per liter, respectively). Accordingly, it is important to assess and quantify the<br />

effects of flowback and produced water on DBP generation (see Section 5.3).<br />

Analytical methods for the measurement of bromide and bromate in elevated TDS matrices are<br />

currently being developed. EPA Method 300.1 is being modified to use a mass spectrometer rather<br />

that an electroconductivity detector, which is unable to detect bromide and bromate in the<br />

presence high anion concentrations (SO 4<br />

2-, NO 2<br />

-, NO 3<br />

-, F - , Cl - ). The mass spectrometer allows<br />

selected ion monitoring specifically for the two natural stable isotopes of bromine ( 79 Br and 81 Br),<br />

with minimal interference from other anions in the high-salt matrix. Interference of the bromide<br />

and bromate response in the mass spectrometer are being assessed by comparing instrument<br />

responses to solutions of bromide and bromate in deionized water with selected anions over a<br />

range of ratios typically encountered in hydraulic fracturing wastewater samples (US EPA, 2012d).<br />

Interference concentration thresholds are being established, and a suitable sample dilution method<br />

is being developed for the quantification of bromide and bromate in actual hydraulic fracturing<br />

wastewater samples. Method detection limits and lowest concentration minimum reporting levels<br />

are being calculated for bromide and bromate in high-salt matrices according to EPA protocols (US<br />

EPA, 2010h).<br />

5.4.4.5. Radionuclides<br />

Gross α and β analyses measure the radioactivity associated with gross α and gross β particles<br />

that are released during the natural decay of radioactive elements, such as uranium, thorium, and<br />

radium. Gross α and β analyses are typically used to screen hydraulic fracturing wastewater in<br />

order to assess gross levels of radioactivity. This information can be used to identify waters<br />

needing radionuclide-specific characterization. The TDS and organic content characteristic of<br />

hydraulic fracturing wastewater, however, interferes with currently accepted methods for gross<br />

α and β analyses. The QAPP for testing and developing gross α and β analytical methods is in<br />

development, and, after it is approved, work will begin.<br />

5.4.4.6. Inorganic Chemicals<br />

In addition to the potential mobilization of naturally occurring radioactive elements, hydraulic<br />

fracturing may also release other elements from the fractured shales, tight sands, and coalbeds,<br />

notably heavy metals such as barium and strontium. Inorganic compounds may also be added to<br />

hydraulic fracturing fluids to perform various functions (e.g., cross-linkers using borate salts, brine<br />

carrier fluids using potassium chloride, and pH-adjusting agents using sodium carbonates) (US EPA,<br />

2011e). Due to the injection or release of naturally occurring metals in unknown quantities, it is<br />

essential that analytical methods for the determination of inorganic elements in waters associated<br />

with hydraulic fracturing be robust and free from interferences that may mask true concentrations.<br />

120


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

The EPA SW-846 Method 6010, employing inductively coupled plasma-optical emission<br />

spectrometry, will be used as a base method for major elements while SW-846 Method 6020 based<br />

on inductively coupled plasma-mass spectrometry will be used as a base method for trace<br />

elements. 70 These methods will be tested and potentially modified for detection of major and trace<br />

elements in hydraulic fracturing wastewater.<br />

5.4.5. Quality Assurance Summary<br />

Three QAPPs have been prepared for the analytical method testing research program. The first<br />

QAPP, “Quality Assurance Project Plan for the Chemical Characterization of Select Constituents<br />

Relevant to Hydraulic Fracturing” (US EPA, 2011g), is the broad general QAPP for the methods<br />

development research project. The QAPP was approved on September 1, 2011. In order to<br />

maintain high QA standards and practices throughout the project, a surveillance audit was<br />

performed on November 15, 2011. The purpose of the surveillance audit was to examine the<br />

processes associated with the in-house extraction of ethoxylated alcohols. Three<br />

recommendations were identified and have been accepted.<br />

The second QAPP, “Formation of Disinfection By-Products from Hydraulic Fracturing Fluid<br />

Constituents Quality Assurance Project Plan,” (US EPA, 2011h), provides details on modifications to<br />

EPA Method 300.1 for optimizing bromide/bromate recoveries in high-salt matrices. The QAPP was<br />

approved on October 5, 2011, and the addendum for bromide/bromate analytic method<br />

development was approved on June 28, 2012. There are no deviations from existing QAPPs to<br />

report at this time. A surveillance audit was performed in March 2011 before the analytical method<br />

addendum (June 28, 2012); therefore, the analytical method development for bromide/bromate<br />

has not yet been audited.<br />

The third QAPP, “Quality Assurance Project Plan for the Inter-Laboratory Verification and<br />

Validation of Diethylene Glycol, Triethylene Glycol, Tetraethylene Glycol, 2-Butoxyethanol and 2­<br />

Methoxyethanol in Ground and Surface Waters by Liquid Chromatography/Tandem Mass<br />

Spectrometry” (US EPA, 2012r), was prepared specifically for the verification of the EPA Region 3<br />

SOP. The QAPP was approved on April 4, 2012. Since then, two surveillance audits and two internal<br />

TSAs have been performed, specifically looking at procedures related to glycol standard<br />

preparation and analysis. The two surveillance audits resulted in one case of potentially mislabeled<br />

samples during stock solution preparation. The potential mislabeling was already identified and<br />

documented by the researchers involved and corrective action taken. The designated EPA QA<br />

Manager found the methods in use satisfactory and further recommendations for improving the QA<br />

process were unnecessary. The internal TSAs also yielded no acts, errors, or omissions that would<br />

have a significant adverse impact on the quality of the final product.<br />

70 Major and trace elements are identified in the retrospective case study QAPPs found at<br />

http://www.epa.gov/<strong>hf</strong>study/qapps.html.<br />

121


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

6. Toxicity Assessment<br />

Throughout the hydraulic fracturing water lifecycle, routes exist through which fracturing fluids<br />

and/or naturally occurring substances could be introduced into drinking water resources. To<br />

support future risk assessments, the EPA is gathering existing data regarding toxicity and potential<br />

human health effects associated with the chemicals reported to be in fracturing fluids and found in<br />

wastewater. At this time, the EPA has not made any judgment about the extent of exposure to these<br />

chemicals when used in hydraulic fracturing fluids or found in hydraulic fracturing wastewater, or<br />

their potential impacts on drinking water resources.<br />

6.1. Relationship to the Hydraulic Fracturing Study<br />

The EPA is compiling existing information on chemical, physical, and toxicological properties of<br />

hydraulic fracturing-related chemicals, which include chemicals reported to be used in injected<br />

hydraulic fracturing fluids and chemicals detected in flowback and produced water. There are<br />

currently over 1,000 chemicals. This work focuses particularly on compiling and evaluating existing<br />

toxicological properties and will inform answers to the research questions listed in Table 47.<br />

Table 47. Secondary research questions addressed by compiling existing information on hydraulic fracturing-related<br />

chemicals.<br />

Water Cycle Stage<br />

Chemical mixing<br />

Flowback and produced water<br />

Applicable Research Questions<br />

What are the chemical, physical, and toxicological properties of<br />

hydraulic fracturing chemical additives<br />

What are the chemical, physical, and toxicological properties of<br />

hydraulic fracturing wastewater constituents<br />

6.2. Project Introduction<br />

Given the potential for accidental human exposure due to spills, improper wastewater treatment,<br />

and potential seepage, it is important to understand the known and potential hazards posed by the<br />

diversity of chemicals needed during hydraulic fracturing. The US House of Representatives’<br />

Committee on Energy and Commerce Minority Staff released a report (2011) noting that more than<br />

650 products (i.e., chemical mixtures) used in hydraulic fracturing contain 29 chemicals that are<br />

either known or possible human carcinogens or are currently regulated under the SDWA (see Table<br />

11 in Section 3.1) (USHR, 2011). However, the report did not characterize the inherent chemical<br />

properties and potential toxicity of many of the reported compounds. The identification of inherent<br />

chemical properties will facilitate the development of models to predict environmental fate,<br />

transport, and the toxicological properties of chemicals. Through this level of understanding,<br />

scientists can design or identify more sustainable alternative chemicals that minimize or even avoid<br />

many fate, transport, and toxicity issues, while maintaining or improving commercial use.<br />

The EPA must understand (1) potential hazards inherent to the chemicals being used in or released<br />

by hydraulic fracturing and returning to the surface in flowback and produced water, (2) doseresponse<br />

characteristics, and (3) potential exposure levels in order to assess the potential impacts<br />

to human health from ingestion of drinking water that might contain the chemicals. The<br />

information from the toxicity assessment project provides a foundation for future risk assessments.<br />

122


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

While the EPA currently does not have plans to conduct a formal risk assessment on this topic, the<br />

information may aid others who are investigating the risk of exposure.<br />

6.3. Research Approach<br />

Once the EPA identifies chemicals reported to be used in hydraulic fracturing fluids or found in<br />

flowback and produced water, physicochemical properties and chemical structures are assigned<br />

using various chemical software packages. Toxicological properties are then identified from<br />

authoritative sources or are estimated based on chemical structure.<br />

Identification of Chemicals. The EPA, to date, has identified nine sources, listed in Table 48, that<br />

contain authoritative information on chemicals in used in hydraulic fracturing fluids or found in<br />

hydraulic fracturing wastewater. The sources have been used to compile two lists: chemicals<br />

reported to be used in hydraulic fracturing fluids and chemicals detected in hydraulic fracturing<br />

wastewater. Chemicals will be added to the two lists as new data become available.<br />

Table 48. References used to develop a consolidated list of chemicals reportedly used in hydraulic fracturing fluids<br />

and/or found in flowback and produced water.<br />

Description / Content<br />

Chemicals reportedly used by 14 hydraulic fracturing service<br />

companies from 2005 to 2009<br />

Products and chemicals used during natural gas operations<br />

with some potential health effects<br />

Chemicals used or proposed for use in hydraulic fracturing<br />

and chemicals found in flowback<br />

Chemicals reportedly used by nine hydraulic fracturing service<br />

companies from 2005 to 2010<br />

MSDSs provided to the EPA during on-site visits<br />

Table 4-1: Characteristics of undiluted chemicals found in<br />

hydraulic fracturing fluids (based on MSDSs)<br />

Chemicals used in Pennsylvania for hydraulic fracturing<br />

activities (compiled from MSDSs)<br />

Chemical records entered in FracFocus for individual wells<br />

from January 1, 2011, through February 27, 2012<br />

Chemicals detected in flowback from 19 hydraulically<br />

fractured shale gas wells in Pennsylvania and West Virginia<br />

Chemicals reportedly detected in flowback and produced<br />

water from 81 wells<br />

Reference<br />

USHR, 2011<br />

Colborn et al., 2011<br />

NYSDEC, 2011<br />

US EPA, 2011b<br />

Material Safety Data Sheets<br />

US EPA, 2004b<br />

PADEP, 2010<br />

GWPC, 2012b<br />

Hayes, 2009<br />

US EPA, 2011k<br />

While compiling the list of chemicals used in fracturing fluids, the EPA identified instances where<br />

various chemical names were reported for a single CASRN. Chemical name and structure<br />

annotation QC methods were applied to the reported chemicals in order to standardize the<br />

chemical names; this process is described in “Chemical Information Quality Review Procedures” for<br />

123


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

the Distributed Structure-Searchable Toxicity (DSSTox) Database Network. 71 The chemical QC<br />

methods included ensuring correct chemical names and CASRNs, and eliminating duplicates where<br />

appropriate. Chemical structures from the DSSTox database were assigned where possible.<br />

Physicochemical Properties. Physicochemical properties of chemicals in the hydraulic fracturing<br />

fluid chemical list were generated from the two-dimensional (2-D) chemical structures from the<br />

EPA’s DSSTox Database Network in structure-data file format. Properties were calculated using<br />

LeadScope chemoinformatic software (Leadscope Inc., 2012), Estimation Programs Interface Suite<br />

for Microsoft Windows (US EPA, 2012a), and QikProp (Schrodinger, 2012). 72 Both Leadscope and<br />

Qikprop software require input of desalted structures. Therefore, the structures were desalted, a<br />

process where salts and complexes are simplified to the neutral, uncomplexed form of the chemical,<br />

using Desalt Batch option in ChemFolder (ACD Labs, 2008). All Leadscope general chemical<br />

descriptors (Parent Molecular Weight, AlogP, Hydrogen Bond Acceptors, Hydrogen Bond Donors,<br />

Lipinski Score, Molecular Weight, Parent Atom Acount, Polar Surface Area, and Rotatable Bonds)<br />

were calculated by default. For EPISuite properties, both the desalted and non-desalted 2-D files<br />

were run using Batch Mode to calculate environmentally relevant, chemical property descriptors.<br />

The chemical descriptors in QikProp require 3-D chemical structures. For these calculations, the 2­<br />

D desalted chemical structures were converted to 3-D using the Rebuild3D function in the<br />

Molecular Operating Environment software (Chemical Computing Group). All computed<br />

physicochemical properties are added into the structure-data file prior to assigning toxicological<br />

properties.<br />

Toxicological Properties. Known and predicted toxicity reference values are being combined into a<br />

single toxicity reference value resource for hydraulic fracturing-related chemicals. The EPA’s list of<br />

hydraulic fracturing-related chemicals was cross-referenced against the following nine sources to<br />

obtain authoritative toxicity reference values:<br />

• US EPA Integrated Risk Information System (IRIS)<br />

• US EPA Provisional Peer-Reviewed Toxicity Value (PPRTV) database<br />

• US EPA Health Effects Assessment Summary Tables<br />

• Agency for Toxic Substances and Disease Registry Minimum Risk Levels<br />

• State of California Toxicity Criteria Database<br />

• State of Alabama Risk-Based Corrective Action document<br />

• State of Florida Cleanup Target Levels<br />

• State of Hawaii Maximum Contaminant List<br />

• State of Texas Effects Screening Levels List<br />

71 For more information on DSSTox, see http://www.epa.gov/ncct/dsstox/ChemicalInfQAProcedures.html.<br />

72 The QikProp, EPI Suite, and LeadScope chemoinformatics programs calculate complementary properties with some<br />

overlap due to the process being performed in batch mode with all default properties included.<br />

124


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Authoritative toxicity reference values have been identified for over 100 of the more than 1,000<br />

chemicals reported as being present in injected water or present in produced water. These include<br />

the benzene, toluene, ethylbenzene, and xylene (BTEX) chemicals, and over 70 others with toxicity<br />

reference values in the IRIS and PPRTV databases.<br />

For the remaining chemicals that lack authoritative toxicity reference values, the structure-data file<br />

(generated for assigning physicochemical properties) can be used with the quantitative structure<br />

toxicity relationship software Toxicity Prediction by Komputer Assisted Technology, or TOPKAT<br />

(Accelrys Discovery Studio, 2012) to identify toxicity values. Rat chronic lowest observed adverse<br />

effect levels (LOAELs) were estimated using the LOAEL module for TOPKAT. The LOAEL module<br />

compares LOAEL values from open literature, National Cancer Institute/National Toxicology<br />

Program technical reports, and EPA databases to estimated rat oral LD 50 values, and then compares<br />

the octanol-water partition coefficient from the chemical structure data file to the range in the<br />

training set.<br />

The estimated LOAEL values will be compared to the authoritative toxicity reference values (for the<br />

chemicals with these authoritative values) to provide an estimate of how similar these values are. It<br />

is important to note that there may be significant deviation between the estimated LOAEL and the<br />

authoritative toxicity reference value for any given chemical due to the use of uncertainty factors in<br />

calculating the reference value, the fact that the reference values are not based on a rat chronic<br />

assay, and whether the reference values are calculated using the benchmark dose, a no observed<br />

adverse effect level, or a LOAEL. However, there is evidence that the estimated LOAEL is generally<br />

within 100 times the concentration of the actual rat chronic LOAEL (Rupp et al., 2010).<br />

6.4. Status and Preliminary Data<br />

Chemicals used in fracturing fluids or found in flowback and produced water, reported by the<br />

sources listed in Table 48, were consolidated and annotated, resulting in lists containing 1,027<br />

unique chemical substances, of which 751 could be assigned a chemical structure and all but 5<br />

assigned CASRNs. Physicochemical properties have been obtained for 318 of the 751 chemicals<br />

with structures. Physicochemical properties for the remainder of the chemicals with structures are<br />

currently being calculated. There were an additional 409 substances that were too poorly defined<br />

in the original lists to be unambiguously designated as unique substances, assigned CASRNs or<br />

chemical structures. The chemical lists are provided in Appendix A. The EPA has completed the first<br />

phase of development for the toxicity reference value database described above.<br />

6.5. Next Steps<br />

The EPA is currently identifying any additional state-based reference value data sources that can be<br />

useful; these additional sources, if any, will be brought into the database as they are identified.<br />

6.6. Quality Assurance Summary<br />

There are two QAPPs associated with this project. The first “Health and Toxicity Theme Hydraulic<br />

Fracturing Study Immediate Office National Center for Environmental Assessment,” was approved<br />

February 2012 and describes the development of the toxicity reference value master spreadsheet<br />

(US EPA, 2012k). The second QAPP, “Health and Toxicity (HT) Hydraulic Fracturing (HF) National<br />

Center for Computational Toxicology,” was approved February 2012 and describes the planning<br />

125


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

and quality processes for the generation of the chemical lists and the calculation of physicochemical<br />

properties for the chemicals for which chemical structures can be assigned (US EPA, 2012i).<br />

126


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

7. Case Studies<br />

7.1. Introduction to Case Studies<br />

Case studies are widely used to conduct in-depth investigations of complex topics and provide a<br />

systematic framework for investigating relationships among relevant factors. In conjunction with<br />

other elements of the research program, they help determine whether hydraulic fracturing can<br />

impact drinking water resources and, if so, the extent and possible causes of any impacts. Case<br />

studies may also provide opportunities to assess the fate and transport of fluids and contaminants<br />

in different regions and geologic settings. Results from the case studies are expected to help answer<br />

the secondary research questions listed in Table 49.<br />

Table 49. Secondary research questions addressed by conducting case studies.<br />

Water Cycle Stage<br />

Chemical mixing<br />

Well injection<br />

Flowback and produced water<br />

Applicable Secondary Research Questions<br />

• If spills occur, how might hydraulic fracturing chemical additives<br />

contaminate drinking water resources<br />

• How effective are current well construction practices at containing<br />

gases and fluids before, during, and after hydraulic fracturing<br />

• Can subsurface migration of fluids or gases to drinking water<br />

resources occur, and what local geologic or man-made features<br />

might allow this<br />

• If spills occur, how might hydraulic fracturing wastewaters<br />

contaminate drinking water resources<br />

Two types of case studies are being conducted as part of this study. Retrospective case studies focus<br />

on investigating reported instances of drinking water resource contamination in areas where<br />

hydraulic fracturing events have already occurred. Prospective case studies involve sites where<br />

hydraulic fracturing will be implemented after the research begins, which allows sampling and<br />

characterization of the site before, during, and after drilling, injection of the fracturing fluid,<br />

flowback, and production. The EPA continues to work with industry partners to design and develop<br />

prospective case studies. Because prospective case studies remain in their early stages, the<br />

progress report focuses on the progress of retrospective case studies only.<br />

To select the retrospective case study sites, the EPA invited stakeholders from across the country to<br />

participate in the identification of locations for potential case studies through informational public<br />

meetings and the submission of electronic or written comments. Following thousands of comments,<br />

over 40 locations were nominated for inclusion in the study. 73 These locations were prioritized and<br />

chosen based on a rigorous set of criteria, including proximity of population and drinking water<br />

supplies, evidence of impaired water quality, health and environmental concerns, and knowledge<br />

gaps that could be filled by a case study at each potential location. Sites were prioritized based on<br />

geographic and geologic diversity, population at risk, geologic and hydrologic features,<br />

characteristics of water resources, and land use (US EPA, 2011e). Five retrospective case study<br />

locations were ultimately chosen for inclusion in this study and are shown in Figure 27.<br />

73 A list of the sites submitted for consideration can be found in the Study Plan.<br />

127


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Figure 27. Locations of the five retrospective case studies chosen for inclusion in the EPA’s Study of the Potential<br />

Impacts of Hydraulic Fracturing on Drinking Water Resources. The locations were nominated by stakeholders and<br />

selected based on criteria described in the text. (ESRI, 2010a, b; US EIA, 2011d, e)<br />

7.1.1. General Research Approach<br />

Although each retrospective case study differs in the geologic and hydrologic characteristics, as well<br />

as the hydraulic fracturing techniques used and the oil and gas exploration and production history<br />

of the area, the methods used to assess potential drinking water impacts are applicable to all of the<br />

study sites. By coordinating the case study methods and analyses, it will be possible to compare the<br />

results of each study. Table 50 describes the general research approach being used for the<br />

retrospective case studies. 74 The tiered scheme uses the results of earlier tiers to refine sampling<br />

activities in later tiers. This approach is both useful and appropriate when the impacts to drinking<br />

water resources and the potential sources of the impacts are unknown. For example, it allows the<br />

sampling to verify key findings and adjust to the improved understanding of the site.<br />

74 The Dunn County, North Dakota, retrospective case study does not use this tiered sampling plan because it is designed<br />

to examine the impacts of a well blowout during hydraulic fracturing. Since the potential source of contamination is<br />

known, the tiered sampling plan is not necessary.<br />

128


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table 50. General approach for conducting retrospective case studies. The tiered approach uses the results of earlier<br />

tiers to refine sampling activities in later tiers.<br />

Tier Goal Critical Path<br />

• Evaluate existing data and information from operators, private<br />

citizens, state and local agencies, and tribes (if any)<br />

1 Verify potential issue<br />

• Conduct site visits<br />

• Interview stakeholders and interested parties<br />

2<br />

3<br />

4<br />

Determine approach<br />

for detailed<br />

investigations<br />

Conduct detailed<br />

investigations to<br />

detect and evaluate<br />

potential sources of<br />

contamination<br />

Determine the<br />

source(s) of any<br />

impacts to drinking<br />

water resources<br />

• Conduct initial sampling of water wells, taps, surface water, and<br />

soils<br />

• Identify potential evidence of drinking water contamination<br />

• Develop conceptual site model describing possible sources and<br />

pathways of the reported or potential contamination<br />

• Develop, calibrate, and test fate and transport model(s)<br />

• Conduct additional sampling of soils, aquifer, surface water, and<br />

wastewater pits/tanks (if present)<br />

• Conduct additional testing, including further water testing with new<br />

monitoring points, soil gas surveys, geophysical testing, well<br />

mechanical integrity testing, and stable isotope analyses<br />

• Refine conceptual site model and further test exposure scenarios<br />

• Refine fate and transport model(s) based on new data<br />

• Develop multiple lines of evidence to determine the source(s) of<br />

impacts to drinking water resources<br />

• Exclude possible sources and pathways of the reported<br />

contamination<br />

• Assess uncertainties associated with conclusions regarding the<br />

source(s) of impacts<br />

Each retrospective case study has developed a QAPP that describes the detailed plan for the<br />

research at that location. The QAPP integrates the technical and quality aspects of the case study in<br />

order to provide a guide for obtaining the type and quality of environmental data required for the<br />

research. Before each new tier of sampling begins, the QAPPs are revised to account for any<br />

changes.<br />

Ground water samples have been collected at all retrospective case study locations. The samples<br />

come from a variety of available sources, such as existing monitoring wells, domestic and municipal<br />

water wells, and springs. Surface water, if present, has also been sampled. During sample collection,<br />

the following water quality parameters were monitored and recorded:<br />

• Temperature<br />

• pH<br />

• TDS<br />

• Specific conductivity<br />

• Alkalinity<br />

• Turbidity<br />

• Dissolved oxygen<br />

129


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

• Oxidation/reduction potential<br />

• Ferrous iron<br />

• Hydrogen sulfide<br />

Each water sample has been analyzed for a suite of chemicals; groups of analytes and examples of<br />

specific chemicals of interest are listed in Table 51. These chemicals include major anions,<br />

components of hydraulic fracturing fluids (i.e., glycols), and potentially mobilized natural occurring<br />

substances (i.e., metals); 75 these chemicals are thought to be present frequently in hydraulic<br />

fracturing fluids or wastewater. As indicated in Table 51, stable isotope analyses are also being<br />

conducted. Stable isotope compositions can be important indicators of what is naturally occurring<br />

in the environment being studied. If an element has multiple stable isotopes, one is usually the most<br />

common form in that environment. Due to different processes that may occur in or around the<br />

environment, other stable isotopes of the element may be found. The different isotopes can make it<br />

easier to determine the source of, or distinguish between, sources of contamination.<br />

Table 51. Analyte groupings and examples of chemicals measured in water samples collected at the retrospective<br />

case study locations.<br />

Analyte Groups<br />

Examples<br />

Anions<br />

Bromide, chloride, sulfate<br />

Carbon group Dissolved organic carbon,* dissolved inorganic carbon †<br />

Dissolved gases<br />

Methane, ethane, propane<br />

Extractable petroleum hydrocarbons Gasoline range organics, § diesel range organics ‡<br />

Glycols<br />

Diethylene glycol, triethylene glycol, tetraethylene glycol<br />

Isotopes<br />

Isotopes of oxygen and hydrogen in water, carbon and<br />

hydrogen in methane, strontium<br />

Low molecular weight acids<br />

Formate, acetate, butyrate<br />

Measures of radioactivity<br />

Radium, gross α, gross β<br />

Metals<br />

Arsenic, manganese, iron<br />

Semivolatile organic compounds Benzoic acid; 1,2,4-trichlorobenzene; 4-nitrophenol<br />

Surfactants<br />

Octylphenol ethoxylate, nonylphenol<br />

Volatile organic compounds<br />

Benzene, toluene, styrene<br />

* Dissolved organic carbon is a result of the decomposition of organic material in aquatic systems.<br />

†<br />

Dissolved inorganic carbon is the sum of the carbonate species (e.g., carbonate, bicarbonate) dissolved in water.<br />

§<br />

Gasoline range organics include hydrocarbon molecules containing 5–12 carbon atoms.<br />

‡<br />

Diesel range organics include hydrocarbon molecules containing 15–18 carbon atoms.<br />

The samples taken for the case studies were analyzed by the EPA Region 8 Laboratory and the EPA<br />

Robert S. Kerr Environmental Research Center. A laboratory TSA was conducted at the EPA Region<br />

8 Laboratory on July 26, 2011; no findings were identified. In addition, a laboratory TSA was<br />

conducted for the onsite analytical support at the Robert S. Kerr Environmental Research Center on<br />

July 28, 2011, which included Shaw Environmental and the EPA General Parameter Lab; no findings<br />

75 A complete list of chemicals and corresponding analytical methods is available in the QAPPs for each case study. See<br />

http://www.epa.gov/<strong>hf</strong>study/qapps.html.<br />

130


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

were identified. The laboratory TSAs were conducted on these laboratories during the first<br />

retrospective case study sampling event to identify any problems early and allow for corrective<br />

actions, if needed. Additional TSAs will be performed if determined to be necessary based on<br />

quality concerns.<br />

This chapter includes progress reports for the following retrospective case studies:<br />

7.2. Las Animas and Huerfano Counties, Colorado ....................................................................................... 131<br />

Investigation of potential drinking water impacts from coalbed methane extraction in the<br />

Raton Basin<br />

7.3. Dunn County, North Dakota ........................................................................................................................... 137<br />

Investigation of potential drinking water impacts from a well blowout during hydraulic<br />

fracturing for oil in the Bakken Shale<br />

7.4. Bradford County, Pennsylvania .................................................................................................................... 142<br />

Investigation of potential drinking water impacts from shale gas development in the Marcellus<br />

Shale<br />

7.5. Washington County, Pennsylvania.............................................................................................................. 148<br />

Investigation of potential drinking water impacts from shale gas development in the Marcellus<br />

Shale<br />

7.6. Wise County, Texas ............................................................................................................................................ 153<br />

Investigation of potential drinking water impacts from shale gas development in the Barnett<br />

Shale<br />

7.2. Las Animas and Huerfano Counties, Colorado<br />

7.2.1. Project Introduction<br />

Las Animas and Huerfano Counties, Colorado, are located on the eastern edge of the Rocky<br />

Mountains and have a combined population of about 22,000 people and a population density of<br />

about 4 people per square mile (USCB, 2010c, d). As shown in Figure 28, the coal-bearing region of<br />

the Raton Basin occupies an area of 1,100 square miles within these two counties. The development<br />

of CBM resources in the Raton and Vermejo Formations within the Raton Basin has increased due<br />

to advances in hydraulic fracturing technology (Keighin, 1995; Watts, 2006b).<br />

131


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Figure 28. Extent of the Raton Basin in southeastern Colorado and northeastern New Mexico (ESRI, 2012; US EIA,<br />

2011d; USCB, 2012a, b, c). The case study includes two locations: “North Fork Ranch,” located northwest of the city<br />

of Trinidad in western Las Animas County, and “Little Creek,” located southwest of the city of Walsenburg in<br />

Huerfano County.<br />

Study site locations in Las Animas and Huerfano Counties were selected in response to ongoing<br />

complaints<br />

about changes<br />

in appearance, odor, and taste associated with drinking water in<br />

domestic wells.<br />

These<br />

sites include “North Fork<br />

Ranch,” located<br />

northwest of the city<br />

of Trinidad in<br />

western<br />

Las Animas County, and<br />

“Little Creek,”<br />

located<br />

southwest<br />

of the<br />

city of<br />

Walsenburg<br />

in<br />

Huerfano<br />

County. In some<br />

locations,<br />

point-of-use water treatment systems have been<br />

installed<br />

on<br />

properties to treat elevated methane and sulfide concentrations in well water. This case<br />

study<br />

focuses on<br />

the potential<br />

impacts of<br />

hydraulic fracturing on<br />

drinking<br />

water resources near<br />

these<br />

two<br />

study sites.<br />

Potential<br />

sources<br />

of ground<br />

water contamination<br />

under consideration<br />

include<br />

activities<br />

associated with natural sources, CBM extraction<br />

( such<br />

as leaking or<br />

abandoned<br />

pits)<br />

, gas<br />

well<br />

completion<br />

and<br />

enhancement techniques, improperly<br />

plugged<br />

and<br />

abandoned<br />

wells, gas<br />

migration,<br />

and residential impacts.<br />

132


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

7.2.2. Site Background<br />

Geology. The Raton Basin is a north-south trending sedimentary and structural depression located<br />

along the eastern edge of the Rocky Mountains, between the Sangre de Cristo Mountains to the west<br />

and the Apishapa, Las Animas, and Sierra Grande arches on the east (Watts, 2006a). This chevronshaped<br />

basin encompasses roughly 2,200 square miles of southeastern Colorado and northeastern<br />

New Mexico, extending from southern Colfax County, New Mexico, through Las Animas County,<br />

Colorado, and northward into Huerfano County, Colorado, as shown in Figure 28 (Tremain, 1980).<br />

It is the southernmost of the several major coal-bearing basins located along the eastern margin of<br />

the Rocky Mountains (Johnson and Finn, 2001). Within the Raton Basin, the Vermejo and Raton<br />

Formations contain CBM resources being extracted using hydraulic fracturing.<br />

Las Animas and Huerfano Counties are underlain by sedimentary bedrock ranging in age from the<br />

late Cretaceous to the Eocene (see Appendix D for a geologic timeline). Igneous intrusions, dating to<br />

the Eocene, Miocene, and Pliocene epochs, occur throughout the area. The sedimentary sequence<br />

exposed within the Raton Basin was deposited in association with regression of the Cretaceous<br />

Interior Seaway, and the stratigraphy reflects deposition in fluvial systems and peat-forming<br />

swamps (Cooper et al., 2007; Flores, 1993). Numerous discontinuous and thin coalbeds are located<br />

in the Vermejo and Raton Formations, which lie directly above the Trinidad Sandstone. The upper<br />

Trinidad intertongues with, and is overlain by, the coal-bearing Vermejo Formation (Topper et al.,<br />

2011). No coal is found below this sandstone (Greg Lewicki & Associates, 2001).<br />

Individual coalbeds in the Vermejo Formation consist of interbedded shales, sandstones, and<br />

coalbeds. The Vermejo Formation ranges in thickness from 150 feet in the southern part of the<br />

basin to 410 feet in the northern part (Greg Lewicki & Associates, 2001). This formation contains<br />

from 3 to 14 coalbeds over 14 inches thick throughout the entire basin, and total coal thickness<br />

typically ranges from 5 to 35 feet (US EPA, 2004b).<br />

The Raton Formation overlies the Vermejo Formation. The Raton Formation ranges from 0 to 2,100<br />

feet thick and is composed of a basal conglomerate, a middle coal-bearing zone, and an upper<br />

transitional zone (Johnson and Finn, 2001; US EPA, 2004b). Its middle coal-bearing zone is<br />

approximately 1,000 feet thick and consists of shales, sandstones, and coalbeds (Greg Lewicki &<br />

Associates, 2001). This zone also contains coal seams that have been mined extensively; total coal<br />

thickness ranges from 10 feet to more than 140 feet in this zone, with individual seams ranging in<br />

thickness from several inches to more than 10 feet (US EPA, 2004b). Sandstones are interbedded<br />

with coalbeds that are currently being developed for CBM, and the coalbeds are the likely source for<br />

gas found in the sandstones (Johnson and Finn, 2001).<br />

Water Resources. Las Animas and Huerfano Counties are located in the Arkansas River Basin and<br />

are drained by the Purgatoire, Apishapa, and Cucharas Rivers. The coal-bearing region of the Raton<br />

Basin is predominantly drained by the Purgatoire and Apishapa Rivers; many stream segments of<br />

these rivers are currently on Colorado’s list of impaired waters (CDPHE, 2012). Annual<br />

precipitation in the Raton Basin is generally correlated to elevation, ranging from over 30 inches<br />

per year in the Spanish Peaks to less than 16 inches per year in eastern portions of the basin, which<br />

are at lower elevation. Much of the precipitation falls as winter snow in the mountains or as intense<br />

summer rain in the plains (Abbott, 1985; S.S. Papadopulos & Associates Inc, 2008). Ground water­<br />

133


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

based drinking water resources in Las Animas and Huerfano Counties reside in four bedrock<br />

aquifers: (1) the Dakota Sandstone and Purgatoire Formation; (2) the Raton Formation, Vermejo<br />

Formation, and Trinidad Sandstone; (3) the Cuchara-Poison Canyon Formation; and (4) volcanic<br />

rocks (Abbott et al., 1983). Sources of recharge to the aquifers include runoff from the Sangre de<br />

Cristo Mountains, precipitation infiltration, and infiltration from streams and lakes (Abbott et al.,<br />

1983; CDM and GBSM, 2004). The depth to ground water depends mostly on topographic position.<br />

In all areas but the southeast corner of the basin, water can be encountered at less than 200 feet<br />

below ground surface (CDM and GBSM, 2004). Regional ground water flow is generally from west<br />

to east, except where it is intercepted by valleys that cut into the rock (Watts, 2006b).<br />

Within the hydrogeologic units of the Raton Basin, sandstone and conglomerate layers transmit<br />

most of the water; shale and coal layers generally retard flow. However, fracture networks in the<br />

shales and coal provide pathways which can transmit fluids or gas. Talus and alluvium may yield<br />

large quantities of water, but are limited in size, and discharges from these units fluctuate<br />

seasonally (Abbott et al., 1983). Aquifer tests in the Raton-Vermejo aquifers indicate hydraulic<br />

conductivities that range from 0 to 45 feet per day (Abbott et al., 1983).<br />

Geologic formations have distinctive ground water chemistry. The Cuchara-Poison Canyon<br />

Formation is typically calcium-bicarbonate type with less than 500 milligrams per liter TDS<br />

content, while the Raton-Vermejo-Trindad aquifer is typically sodium-bicarbonate with TDS<br />

concentrations less than 1,500 milligrams per liter. Abbott et al. (1983) note that concentrations of<br />

boron, fluoride, iron, manganese, mercury, nitrate, selenium, and zinc are locally elevated due to a<br />

variety of geologic processes and human activities. High concentrations of fluoride occur in the<br />

Poison Canyon and Raton Formations, possibly due to the dissolution of detrital fluorite. Iron and<br />

manganese concentrations may be also elevated, particularly in areas where coals are present, due<br />

to the dissolution of pyrite and/or siderite contained in the coal seams. Nitrate enrichment often<br />

occurs in alluvial aquifers where fertilizers or animal wastes add nitrogen (Abbott et al., 1983).<br />

Oil and Gas Exploration and Production. The Raton Basin contains substantial amounts of high- and<br />

medium-volatile bituminous coals, which extend from outcrops along the periphery of the region to<br />

depths of at least 3,000 feet in the deepest parts of the region (Jurich and Adams, 1984). Most of<br />

these coal resources are in the Vermejo and Raton Formations, which are the target formations for<br />

CBM production (Macartney, 2011; Tyler, 1995). These coalbeds have been extensively mined in<br />

the peripheral outcrop belt along major stream valleys, as well as in a few structural uplifts within<br />

the interior of the basin (Dolly and Meissner, 1977). Total coal resources estimated in the basin<br />

range from 1.5 billion to more than 17 billion short tons (Flores and Bader, 1999).<br />

Production of natural gas in the Raton Basin began in the 1980s, but before 1995, there were no gas<br />

distribution lines out of the basin and fewer than 60 wells had been drilled (S.S. Papadopulos &<br />

Associates Inc, 2008). The Raton Basin is estimated to contain as much as 18.4 trillion cubic feet of<br />

CBM (Tyler, 1995). This area has recently seen a rapid expansion in the production of natural gas<br />

with recent advances in hydraulic fracturing technology. Between 1999 and 2004, annual<br />

production of Raton Basin CBM in Las Animas and Huerfano Counties increased from about 28<br />

billion cubic feet to about 80 billion cubic feet, and the number of producing wells grew from 478<br />

wells to 1,543 wells. During the same period, annual ground water withdrawals for CBM production<br />

134


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

increased from about 1.45 billion gallons to about 3.64 billion gallons (Watts, 2006b). Expansion of<br />

CBM wells has focused on the development of the Vermejo coals, since these coals are thicker and<br />

more continuous than those located in the Raton Formation (US EPA, 2004b).<br />

7.2.3. Research Approach<br />

A detailed description of the sampling methods and procedures for this case study can be found in<br />

the project’s QAPP (US EPA, 2012o). Ground water and surface water sampling in this area is<br />

intended to provide a survey of water quality in Las Animas and Huerfano Counties. Data collection<br />

involves sampling water from domestic wells, surface water bodies (streams), monitoring wells, 76<br />

and gas production wells at locations in both Las Animas and Huerfano Counties, as indicated in<br />

Figure 29. The locations of these sampling sites were chosen based on their proximity to production<br />

activity.<br />

Figure 29. Locations of sampling sites in Las Animas and Huerfano Counties, Colorado. Water samples have been<br />

taken from domestic wells, surface water bodies (streams), monitoring wells, and gas production wells.<br />

In addition to the analytes discussed in Section 7.1.1, the stable isotope compositions of carbon and<br />

hydrogen in methane, as well as the stable carbon isotope composition of dissolved inorganic<br />

carbon and the stable sulfur isotope composition of dissolved sulfate and dissolved sulfide, are<br />

76 Monitoring wells were installed by either Pioneer Natural Resources or Petroglyph Energy.<br />

135


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

being analyzed as part of this case study. Microbial analyses are also being conducted on water<br />

samples collected at this case study location in order to better understand the biogeochemical<br />

cycling of carbon and sulfur in ground water. Together, these measurements support the objective<br />

of determining if ground water resources have been impacted, and, if so, whether they were<br />

impacted by hydraulic fracturing activities or other sources of contamination.<br />

7.2.4. Status and Preliminary Data<br />

As of August 2012, two sampling trips have been conducted: one in October 2011 and another in<br />

May 2012. During the October 2011 sampling trip, two production wells, five monitoring wells, 14<br />

domestic water wells, and one surface water location were sampled. During the May 2012 sampling<br />

trip, two production wells, three monitoring wells, 12 domestic water wells, and three surface<br />

water locations were sampled. The locations of sampling sites are displayed in Figure 29.<br />

7.2.5. Next Steps<br />

Additional fieldwork to collect ground and surface water at each sampling location is tentatively<br />

scheduled for late 2012 and spring 2013. Sampling locations and analytes measured may be refined<br />

based on the results of the first two sets of samples. More focused investigations will also be<br />

conducted, if warranted, at locations where potential impacts associated with hydraulic fracturing<br />

may have occurred.<br />

7.2.6. Quality Assurance Summary<br />

The initial QAPP for this case study, “Hydraulic Fracturing Retrospective Case Study, Raton Basin,<br />

CO,” was approved by the designated EPA QA Manager on September 20, 2011 (US EPA, 2012o). A<br />

revision to the QAPP was made before the second sampling event and was approved on April 30,<br />

2012, to update project organization, update lab accreditation information, update sampling<br />

methodology, add sulfur isotope analyses, modify critical analytes, and change the analytical<br />

method for determining hydrogen and oxygen stable isotope ratios in water . There have been no<br />

significant deviations from the QAPP during any sampling event, and therefore no impact on data<br />

quality. A field TSA was conducted on October 4, 2011, during the first sample collection event; no<br />

findings were identified. See Section 7.1.1 for information related to the laboratory TSAs.<br />

As results are reported and raw data are provided from the laboratories, ADQs are performed to<br />

verify that the quality requirements specified in the approved QAPP were met. Data will be<br />

qualified, if necessary, based on these ADQs. The results of these ADQs will be reported in the final<br />

report on this project.<br />

136


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

7.3. Dunn County, North Dakota<br />

7.3.1. Project Introduction<br />

Dunn County, North Dakota, is a rural county with a population of 3,500 and an average population<br />

density of 1.8 people per square mile (USCB, 2010b); Killdeer is its largest city. This part of North<br />

Dakota is currently experiencing renewed natural gas exploration and a boom in oil production<br />

from the Bakken Shale, which extends domestically from western North Dakota to parts of<br />

northeastern Montana (Figure 30). The area’s increased oil and gas exploration has relied greatly<br />

upon both horizontal drilling and hydraulic fracturing technologies.<br />

Figure 30. Extent of the Bakken Shale in North Dakota and Montana (US EIA, 2011d; USCB, 2012a, c). The case<br />

study focuses on a well blowout that occurred in Dunn County, North Dakota, in September 2010.<br />

The EPA’s case study site in Killdeer, North Dakota, was chosen at the request of the state to<br />

specifically examine any water resource impacts from a well blowout in September 2010 that<br />

resulted in an uncontrolled release of hydraulic fracturing fluids and formation fluids. The Killdeer<br />

Aquifer, the main source of drinking water for the city of Killdeer, underlies the study site. The<br />

137


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

blowout occurred at the Franchuk 44-20 SWH well, which is just outside the Killdeer municipal<br />

water supply well’s 2.5 mile wellhead protection zone.<br />

The uncontrolled blowout occurred on September 1, 2010, during the fifth stage of a hydraulic<br />

fracturing treatment of the Franchuk 44-20 SWH well. The intermediate well casing burst because<br />

of a 8,390 pounds per square inch pressure spike that released the pop-off relief valve. Hydraulic<br />

fracturing fluids and formation fluids began flowing from the ground around the well at several<br />

points and then flowed toward the northeast corner of the well pad, where they were contained by<br />

a 2 foot berm. During that day, 47,544 gallons of fluids were removed from the site. The following<br />

day, 88,000 gallons of fluids were removed from the site, and 15,120 gallons of mud were circulated<br />

into the well to kill it. Three monitoring wells were installed, but not sampled. Two down-gradient<br />

homeowner wells, an up-gradient homeowner well, and two municipal water wells were sampled<br />

on September 2. Three cement plugs were installed beginning at 9,000 feet in the wellbore, and<br />

105,252 gallons of fluid were removed from the site. A bridge plug was set at 9,969 feet on<br />

September 6. From September 30 to October 15, 2,000 tons of contaminated soil were removed and<br />

disposed of (Jacob, 2011). Since the blowout, the State of North Dakota has overseen site cleanup<br />

and has required the well’s operator to conduct ground water monitoring on a quarterly basis. In<br />

November 2010, the state asked the EPA to consider this site as part of this study, and the EPA<br />

agreed to do so.<br />

7.3.2. Site Background<br />

Geology. Dunn County is located in west-central North Dakota and is underlain by the sedimentary<br />

rocks of the Williston Basin. Although Dunn County marks the southern extent of glaciations in<br />

North Dakota, most of the glacial deposits have been eroded and the surface sediments are<br />

characterized by post-glacial, channel-fill deposits (Murphy, 2001). As described in Nordeng<br />

(2010), the Bakken formation is primarily composed of shale and dolomite, with some sandstone<br />

and siltstone. The Bakken Shale is of Late Devionian-Early Mississippian age (Appendix D) and is an<br />

organic-rich marine shale. It has no surface outcrop and is constrained by the Madison Formation<br />

above and the Wabamum, Big Valley, and Torquary Formations below (Murphy, 2001; Nordeng,<br />

2010). The depths to the Bakken Shale range from 9,500 to 10,500 feet and its thickness ranges<br />

from very thin up to 140 feet (Carlson, 1985; Murphy, 2001).<br />

Water Resources. Dunn County is a semi-arid region. Surface water in Dunn County is in the<br />

Missouri River Basin and includes the Little Missouri River to the northwest of the county and Lake<br />

Sakakawea to the northeast. These water resources supply water for domestic use, irrigation,<br />

industrial water, and hydraulic fracturing.<br />

One of the major sources of drinking water in Killdeer is the Killdeer Aquifer: a glacial outwash<br />

aquifer, composed of fine to medium sand with course gravel near its base. It is shallow, with a<br />

maximum thickness of 233 feet. The aquifer is generally overlain by clay and silt soils (Klausing,<br />

1979). Yields from the Killdeer Aquifer are high, ranging from 50 to 1,000 gallons per minute<br />

(Klausing, 1979). The major water types in the Killdeer Aquifer are sodium bicarbonate and sodium<br />

sulfate. Table 52 shows background water quality data for the Killdeer Aquifer, compiled by<br />

Klausing (1979).<br />

138


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table 52. Background water quality data for the Killdeer Aquifer in North Dakota (Klausing, 1979). The range of<br />

boron, chloride, and iron in some samples was below the detection limit (BDL).<br />

Parameter<br />

Bicarbonate<br />

Boron<br />

Chloride<br />

Fluoride<br />

Iron<br />

Nitrate<br />

Sodium<br />

Sulfate<br />

TDS<br />

Concentration Range<br />

(milligrams per liter)<br />

374–1,250<br />

BDL–3.70<br />

BDL–25<br />

0.1–2<br />

BDL–5.50<br />

0.3–6.7<br />

50–1,350<br />

333–3,000<br />

234–5,030<br />

Mean Concentration<br />

(milligrams per liter)<br />

713<br />

0.53<br />

4.5<br />

0.66<br />

1.03<br />

1.2<br />

413<br />

626<br />

1,531<br />

Oil and Gas Exploration and Production. Although it was known to contain large volumes of oil as<br />

early as the 1950s, difficulties in extracting the oil from the Bakken Shale kept production rates low<br />

(NDIC, 2012a). Hydraulic fracturing and horizontal drilling technologies have created greater<br />

access to the Bakken Shale oil reserves. In January 2003, Dunn County had 99 wells, producing<br />

approximately 86,000 barrels of oil (NDIC, 2003). By July 2012, the county had 854 wells,<br />

producing approximately 3.2 million barrels of oil (NDIC, 2012b).<br />

7.3.3. Research Approach<br />

A detailed description of this case study’s sampling methods and procedures can be found in the<br />

QAPP (US EPA, 2011i). The primary objective of this case study is to assess the impacts of the<br />

Franchuk 44-20 SWH well blowout that occurred on September 1, 2010. Unlike the EPA’s other four<br />

retrospective case studies, the Killdeer case study does not use a tiered approach because the<br />

potential source of contamination is known. Ground water sampling includes domestic, municipal,<br />

water supply, and monitoring wells. 77 Figure 31 shows the sampling locations in Dunn County,<br />

North Dakota.<br />

77 Terracon Consultants was contracted by the well operator, Denbury Resources, for the installation of monitoring wells.<br />

139


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Figure 31. Location of sampling sites in Dunn County, North Dakota.<br />

Domestic, municipal, and supply wells are being sampled at a tap as close to the wellhead as<br />

possible, before any treatment has occurred. Monitoring wells have been installed and have<br />

dedicated bladder pumps for sampling and purging operations. Water samples collected at these<br />

locations are being analyzed for the chemicals listed in Section 7.1.1 as well as the chemicals listed<br />

in the QAPP (US EPA, 2011i). The data collected as part of this case study will be compared to<br />

existing background data as part of the initial screening phase (Tier 2 in Table 50) to determine if<br />

any contamination has occurred in the study location.<br />

7.3.4. Status and Preliminary Data<br />

Two rounds of sampling were conducted in Killdeer in July and October 2011. Samples were<br />

collected at 10 monitoring wells, three domestic water wells, two water supply wells, and one<br />

municipal water well. The locations of sampling sites are displayed in Figure 31.<br />

7.3.5. Next Steps<br />

At least one more round of sampling is planned to verify data collected from the first two rounds of<br />

sampling. Additional sampling locations or analytes may be included in future rounds as analytical<br />

data are evaluated and additional pertinent information becomes available.<br />

140


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

7.3.6. Quality Assurance Summary<br />

The initial QAPP for this case study, “Hydraulic Fracturing Retrospective Case Study, Bakken Shale,<br />

Killdeer and Dunn County,” was approved by the designated EPA QA Manager on June 20, 2011 (US<br />

EPA, 2011i). A revision to the QAPP was made before the second sampling event and was approved<br />

on August 31, 2011, to address the collection of isotopic samples; revised sampling protocols for<br />

domestic, supply, and municipal wells; and analytical lab information. Another QAPP revision has<br />

been submitted for review by QA staff in preparation for the third sampling event. There have been<br />

no significant deviations from the QAPPs during earlier sampling events, and therefore no impact to<br />

data quality. A field TSA was conducted on July 19, 2011; no findings were identified. See Section<br />

7.1.1 for information related to the laboratory TSAs.<br />

As results are reported and raw data are provided from the laboratories, ADQs will be performed to<br />

verify that the quality requirements specified in the approved QAPP were met. Data will be<br />

qualified if necessary, based on these ADQs. The results of these ADQs will be reported in the final<br />

report on this project.<br />

141


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

7.4. Bradford County, Pennsylvania<br />

7. 4. 1.<br />

Project<br />

Introduction<br />

Bradford<br />

County is<br />

a rural<br />

county in northeastern Pennsylvania<br />

with an approximate total<br />

population<br />

of 63,000<br />

and<br />

an average population<br />

density of 55<br />

people per square mile<br />

( USCB,<br />

2010a). As shown in Figure 32, the Marcellus Shale underlies Bradford County, extending<br />

through<br />

much of New York, Pennsylvania, Ohio, and West Virginia. Recently, natural gas drilling in the<br />

Marcellus Shale has increased significantly in northeastern Pennsylvania, including Bradford<br />

County.<br />

Figure 32. Extent of the Marcellus Shale, which underlies large portions of New York, Ohio, Pennsylvania, and West<br />

Virginia ( US EIA, 2011d; USCB, 2012a, c). The case study focuses on reported changes in drinking water quality in<br />

Bradford County, Pennsylvania, with a few water samples taken in neighboring Susquehanna County.<br />

142


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

The EPA chose Bradford County, and parts of neighboring Susquehanna County, 78 as a retrospective<br />

case study location because of the extensive hydraulic fracturing activities occurring there,<br />

coincident with the large number of homeowner complaints regarding the appearance, odor, and<br />

possible health impacts associated with water from domestic wells. Additionally, the Pennsylvania<br />

Department of Environmental Protection has issued notices of violation for infractions at wells in<br />

this area, including a gas well blowout in Leroy Township of Bradford County in April 2011 that<br />

released a reported 10,000 gallons of flowback and produced water (SAIC Energy Environment &<br />

Infrastructure LLC and Groundwater & Environmental Services Inc., 2011). Initial sampling<br />

locations for this retrospective case study were chosen primarily based on individual homeowner<br />

complaints or concerns regarding potential adverse impacts to their well water from nearby<br />

hydraulic fracturing activities. If anomalies in ground water quality are found during sampling, all<br />

potential sources of contamination in the study area will be considered, including those not related<br />

to hydraulic fracturing.<br />

7.4.2. Site Background<br />

Geology. The geology of the study area has been extensively described in other studies and is<br />

summarized below (Carter and Harper, 2002; Milici and Swezey, 2006; Taylor, 1984; Williams et al.,<br />

1998). The Bradford County study area is underlain by unconsolidated deposits of glacial and postglacial<br />

origin and the nearly flat-lying sedimentary bedrock of the Appalachian Basin. The glacial<br />

and post-glacial deposits consist of till, stratified drift, alluvium, and swamp deposits. The bedrock<br />

consists primarily of shale, siltstone, and sandstone of Devonian to Pennsylvanian age. The<br />

Devonian bedrock includes the Loch Haven and Catskill formations, both of which are important<br />

sources of drinking water in the study area. The Marcellus Shale, also known as the Marcellus<br />

Formation, is a Middle Devonian-age (Appendix D) shale with a black color, low density, and high<br />

organic carbon content. It occurs in the subsurface beneath much of Ohio, West Virginia,<br />

Pennsylvania, and New York (Figure 32). Smaller areas of Maryland, Kentucky, Tennessee, and<br />

Virginia are also underlain by the Marcellus Shale. In Bradford County, the Marcellus Shale<br />

generally lies 4,000 to 7,000 feet below the surface and ranges in thickness from 150 to 300 feet<br />

(Marcellus Center for Outreach and Research, 2012a, b). The Marcellus Shale is part of a<br />

transgressive sedimentary package, formed by the deposition of terrestrial and marine material in a<br />

shallow, inland sea. It is underlain by the sandstones and siltstones of the Onondaga Formation and<br />

overlain by the carbonate rocks of the Mahantango Formation.<br />

Within the Marcellus Shale, natural gas occurs within the pore spaces of the shale, within vertical<br />

fractures or joints of the shale, and adsorbed onto mineral grains and organic material. An<br />

assessment conducted by the USGS in 2011 suggested that the Marcellus Shale contains an<br />

estimated 84 trillion cubic feet of technically recoverable natural gas (Coleman et al., 2011).<br />

78 Four wells were sampled in Susquehanna County during the first round of sampling. Soon after, EPA Region 3 began an<br />

investigation of potential drinking water contamination in Dimock, located in Susquehanna County (see<br />

http://www.epa.gov/aboutepa/states/pa.html). In order to avoid duplication of effort, this case study focuses on<br />

reported changes in drinking water quality in Bradford County. Subsequent sampling for this case study has been, and<br />

will continue to be, done in Bradford County.<br />

143


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Water Resources. The average precipitation in Bradford County is 33 inches per year. Summer<br />

storms produce about half of this precipitation; the remainder of the precipitation, and much of the<br />

ground water recharge, occurs during winter and spring (PADEP, 2012). Surface water in the study<br />

area is part of the Upper Susquehanna River Basin. The main branches of the Susquehanna River<br />

flow to the south, while the smaller tributaries are constrained by the northeast-southwest<br />

orientation of the Appalachian Mountains. Stratified drift aquifers and the Loch Haven and Catskill<br />

bedrock formations serve as primary ground water drinking sources in the study area. Glacial till is<br />

also tapped as a drinking water source at some locations (Williams et al., 1998). These resources<br />

provide water for domestic use, municipal water, manufacturing, irrigation, and hydraulic<br />

fracturing.<br />

The stratified drift aquifers in Bradford County occur as either confined or unconfined aquifers. The<br />

confined aquifers in the study area are composed of sand and gravel deposits of glacial, ice-contact<br />

origin and are typically buried by pro-glacial lake deposits; the unconfined aquifers are composed<br />

of sand and gravel deposited by glacial outwash or melt-waters. Depth to ground water varies<br />

throughout Bradford County and ranged from 1 to 300 feet for the wells sampled in the study. The<br />

median specific capacity of confined stratified drift aquifers is 11 gallons per minute per foot; the<br />

median specific capacity of unconfined stratified drift aquifers is 24 gallons per minute per foot<br />

(Williams et al., 1998). The specific capacity of wells completed in till or bedrock is typically 10<br />

times lower than in the stratified drift aquifers.<br />

Ground water in the study area is generally of two types: a calcium bicarbonate type in zones of<br />

unconfined flow and a sodium chloride type in zones of confined flow. Data from Williams et al.<br />

(1998) show that water wells completed in zones with more confined flow contain higher TDS<br />

(median concentration of 830 milligrams per liter), dissolved barium (median concentration of 2.0<br />

milligrams per liter), and dissolved chloride (median concentration of 349 milligrams per liter)<br />

compared to zones with unconfined flow. This is also true for concentrations of iron and manganese<br />

in the study area. Table 53 presents a summary of median and maximum concentrations of<br />

inorganic parameters in Bradford County ground water, based on the study conducted by Williams<br />

et al. (1998).<br />

Table 53. Background (pre-drill) water quality data for ground water wells in Bradford County, Pennsylvania (Williams<br />

et al., 1998).<br />

Parameter<br />

Median Concentration<br />

(milligrams per liter)<br />

Pre - Drill Data<br />

Maximum Concentration<br />

(milligrams per liter)<br />

Number of<br />

Samples<br />

Arsenic 0.009 0.072 16<br />

Barium 0.175 98 50<br />

Chloride 11 3,500 93<br />

Iron 0.320 15.9 95<br />

Manganese 0.120 1.03 77<br />

TDS 246 6,100 102<br />

pH (pH units) 7.25 8.8 102<br />

144


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Naturally high levels of TDS, barium, and chloride found in ground water make it difficult to assess<br />

the potential impacts of hydraulic fracturing activities in this part of the country since these<br />

analytes would normally serve as indicators of potential impacts. In addition, methane occurs<br />

naturally in ground water in the study area, making an assessment of potential impacts of methane<br />

due to hydraulic fracturing on drinking water resources more challenging than at other study<br />

locations.<br />

Oil and Gas Exploration and Production. Gas drilling to depths of the Marcellus Shale and beyond<br />

dates back to the 1930s, although at that time, the Marcellus Shale was of little interest as a source<br />

of gas. Instead, gas was sought primarily from sandstone and limestone deposits, and the Marcellus<br />

Shale was only encountered during drilling to deeper targeted zones like the Oriskany Sandstone.<br />

Upon penetrating the Marcellus Shale, significant but generally short-lived gas flow would be<br />

observed. With the advent of modern hydraulic fracturing technology and the increasing price of<br />

gas, the Marcellus Shale has become an economical source of natural gas with the potential to<br />

produce several hundred trillion cubic feet (Milici and Swezey, 2006). In July 2008, there were only<br />

48 active permitted natural gas wells in Bradford County; by January 2012, there were 2,015<br />

(Bradford County Government, 2012). The wells are located throughout the county with an average<br />

density of actively permitted wells of 1.8 wells per square mile.<br />

7.4.3. Research Approach<br />

Methods for sampling ground water and surface water are described in detail in the QAPP (US EPA,<br />

2012m). The primary objective of this case study is to determine if ground water resources have<br />

been impacted, and whether or not those impacts were caused by hydraulic fracturing activities or<br />

other sources. Water samples have been taken from domestic wells, springs, ponds, and streams<br />

near gas well pads. Figure 33 shows the sampling locations, which were primarily chosen based on<br />

individual homeowner complaints or concerns regarding potential adverse impacts to water<br />

resources from nearby hydraulic fracturing activities.<br />

145


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Figure 33. Location of sampling sites in Bradford and Susquehanna Counties, Pennsylvania. Samples were taken in<br />

Susquehanna County during the first round of sampling. Later rounds of sampling are focused only in Bradford<br />

County.<br />

In addition to the analytes described in Section 7.1.1, the stable isotope compositions of carbon and<br />

hydrogen in dissolved methane and of carbon in dissolved inorganic carbon are being measured to<br />

determine the potential origin of the methane (i.e., biogenic versus thermogenic). 79 Since methane<br />

is known to be naturally present in the ground water of northeastern Pennsylvania, it is critical to<br />

understand the origin of any methane detected as part of this case study. Samples are also being<br />

analyzed for radium-226, radium-228, and gross alpha and beta radiation, as they may be potential<br />

indicators of hydraulic fracturing impacts to ground water in northeast Pennsylvania. Together,<br />

these measurements support the objective of determining if ground water resources have been<br />

impacted by hydraulic fracturing activities or other sources of contamination.<br />

7.4.4. Status and Preliminary Data<br />

Two rounds of sampling have been completed from 34 domestic wells, two springs, one pond, and<br />

one stream. The first sampling round was conducted in October and November of 2011 and the<br />

second round in April and May of 2012. The locations of sampling sites are displayed in Figure 33.<br />

79 Biogenic methane is formed as methane-producing microorganisms chemically break down organic material.<br />

Thermogenic methane results from the geologic formation of fossil fuel.<br />

146


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

7.4.5. Next Steps<br />

A third round of sampling to verify data collected from the first two rounds of sampling is already<br />

planned. Additional sampling locations may be included and there may be future rounds of<br />

sampling as analytical data from the first three rounds are evaluated and additional pertinent<br />

information becomes available. More focused investigations may also be conducted, if warranted, at<br />

locations where potential impacts associated with hydraulic fracturing are suspected.<br />

7.4.6. Quality Assurance Summary<br />

The initial QAPP for this case study, “Hydraulic Fracturing Retrospective Case Study, Bradford-<br />

Susquehanna Couties, PA,” was approved by the designated EPA QA Manager on October 3, 2011<br />

(US EPA, 2012m). A revision to the QAPP was made prior to the second sampling event and was<br />

approved on April 12, 2012, to address the addition of analytes such as radium-226, radium-228,<br />

lithium, and thorium; updated project organization and accreditation information; and clarification<br />

on some sampling and laboratory QA/QC issues. There have been no significant deviations from the<br />

QAPP during any sampling event, and therefore no impact to data quality. A field TSA was<br />

conducted on October 27, 2011; no findings were identified. See Section 7.1.1 for information<br />

related to the laboratory TSAs.<br />

As results are reported and raw data are provided from the laboratories, ADQs are performed to<br />

verify that the quality requirements specified in the approved QAPP were met. Data will be<br />

qualified if necessary, based on these ADQs. The results of these ADQs will be reported in the final<br />

report on this project.<br />

147


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

7.5. Washington County, Pennsylvania<br />

7.5.1. Project Introduction<br />

Washington County, located about 30 miles southwest of Pittsburgh, Pennsylvania, has a population<br />

of about 208,000 with approximately 240 people per square mile (USCB, 2010e). Figure 34 shows<br />

its position in the western region of the Marcellus Shale. Recently, oil and gas exploration and<br />

production in this area have increased, primarily due to production of natural gas from the<br />

Marcellus Shale using hydraulic fracturing.<br />

Figure 34. Extent of the Marcellus Shale, which underlies large portions of New York, Ohio, Pennsylvania, and West<br />

Virginia (US EIA, 2011d; USCB, 2012a, c). The case study focuses on reported changes in drinking water quality and<br />

quantity in Washington County, Pennsylvania.<br />

The location of this case study was chosen in response to homeowner complaints about changes to<br />

water quality and water quantity in Washington County. Residents in several areas of Washington<br />

County have reported impacts to their private drinking water wells, specifically increased turbidity,<br />

discoloration of sinks, and transient organic odors. Sampling locations were selected in May 2011<br />

by interviewing individuals about their water quality and the timing of any possible water quality<br />

148


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

changes in relation to gas production activities. Potential sources of ground water and surface<br />

water contamination under consideration at this case study site may include activities associated<br />

with oil and gas production (such as leaking or abandoned pits), gas well completion and<br />

enhancement techniques, and improperly plugged and abandoned wells, as well as activities<br />

associated with residential and agricultural practices.<br />

7.5.2. Site Background<br />

Geology. Washington County, like Bradford County, is located in the Appalachian Basin. The geology<br />

of this area of Pennsylvania consists of thick sequences of Paleozoic Era (Appendix D) sedimentary<br />

formations that dip and thicken to the southeast toward the basin axis. The surface geology in<br />

Washington County consists of Quaternary alluvial deposits, predominantly in stream valleys of the<br />

county. Alluvial deposits are generally less than 60 feet thick and consist of clay, silt, sand, and<br />

gravel derived from local bedrock. The formations of the Appalachian Basin are derived from a<br />

variety of clastic and biochemical sedimentary deposits, ranging from terrestrial swamps to nearshore<br />

environments and deep marine basins, which created shales, limestones, sandstones,<br />

coalbeds, and other sedimentary rocks (Shultz, 1999). As previously noted, the Marcellus Shale<br />

formation is of particular importance to recent gas exploration and production in the Appalachian<br />

Basin. In Washington County, the depth to the Marcellus Shale ranges from about 5,000 to 7,000<br />

feet below ground surface (Marcellus Center for Outreach and Research, 2012a). The thickness of<br />

the Marcellus Shale in Washington County is less than 150 feet (Marcellus Center for Outreach and<br />

Research, 2012b).<br />

Water Resources. The rivers and streams of Washington County drain into the Ohio River to the<br />

west. Drinking water aquifers in the county exist in both the alluvial deposits overlying bedrock in<br />

the stream valleys and in the bedrock. Ground water flow in the shallow aquifer system generally<br />

follows the topography, moving from recharge areas near hilltops to discharge areas in valleys.<br />

Background information on the geology and hydrology of Washington County is summarized from<br />

reports published by Newport (1973) and Williams et al. (1993). Ground water in Washington<br />

County occurs in both confined and unconfined aquifers, with well yields ranging from a fraction of<br />

a gallon per minute to over 350 gallons per minute. In this area, water-bearing zones are generally<br />

no deeper than 150 feet below ground surface, and the depth to water varies from 20 to 60 feet<br />

below land surface depending on topographic setting. In addition to alluvial aquifers, ground water<br />

is derived from bedrock aquifers, including the Monongahela Group, the Conemaugh Group, and the<br />

Greene and Washington formations, which consist of limestones, shales, and sandstone units. In<br />

general, ground water derived from these formations has yields ranging from less than 1 to over 70<br />

gallons per minute, and the formations range in depth from less than 40 feet to over 400 feet. The<br />

Conemaugh Group generally provides the greatest yield; the median yield for wells in this aquifer is<br />

5 gallons per minute.<br />

The quality of ground water in Washington County is variable and depends on factors such as<br />

formation lithology and residence time. For example, recharge ground water sampled from hilltops<br />

and hillsides is typically calcium-bicarbonate type and usually low in TDS (about 500 milligrams<br />

per liter). Ground water from valley settings in areas of discharge is typically sodium-bicarbonate<br />

or sodium-chloride type, with higher TDS values (up to 2,000 milligrams per liter). Williams et al.<br />

149


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

(1993) report that background concentrations of iron and manganese in the ground water from<br />

Washington County are frequently above the EPA’s secondary MCLs: over 33% of water samples<br />

had iron concentrations greater than 0.3 milligrams per liter, and 30% of water samples had<br />

manganese concentrations above 0.05 milligrams per liter. Hard water was also reported as being a<br />

common problem in the county, with TDS levels in more than one-third of the wells sampled by<br />

Williams et al. (1993) exceeding 500 milligrams per liter. Arsenic, cadmium, chromium, copper,<br />

lead, mercury, selenium, silver, and zinc were also detected at low levels. Historically, ground water<br />

quality in Washington County has been altered due to drainage from coal mining operations<br />

(Newport, 1973). Additionally, fresh water aquifers in some locations have been contaminated by<br />

brine from deeper non-potable aquifers through historic oil and gas wells that were improperly<br />

abandoned or have corroded casings (Newport, 1973).<br />

Oil and Gas Exploration and Production. The oil and gas development in Washington County dates<br />

back to the 1800s, but generally did not target the Marcellus Shale (Ashley and Robinson, 1922).<br />

The first test gas well into the Marcellus Shale was drilled in Mount Pleasant Township in<br />

Washington County in 2003 and was hydraulically fractured in 2004. Data provided by the<br />

Pennsylvania Department of Environmental Protection indicate that the number of permitted gas<br />

wells in the Washington County area of the Marcellus Shale increased rapidly, from 10 wells in<br />

2005 to 205 wells in 2009 (MarcellusGas.Org, 2012b). From 2009 to 2012, the number of newly<br />

permitted wells per year has remained below 240 (MarcellusGas.Org, 2012c). The anticipated<br />

water usage for all permitted wells in Washington County is estimated to be nearly 5 billion gallons<br />

(MarcellusGas.Org, 2012a).<br />

7.5.3. Research Approach<br />

Methods for sampling ground water and surface water are described in detail in the QAPP (US EPA,<br />

2012n). Samples have been taken from domestic wells and surface water bodies. The EPA chose<br />

sampling locations by interviewing individuals about their water quality and the timing of water<br />

quality changes in relation to gas production activities. The locations of sampling sites are shown in<br />

Figure 35.<br />

150


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Figure 35. Sampling locations in Washington County, Pennsylvania.<br />

Water samples collected at these locations are being analyzed for the chemicals listed in Section<br />

7.1.1 as well as the chemicals listed in the QAPP (US EPA, 2012n). Together these measurements<br />

support the objective of determining if ground water resources have been impacted by hydraulic<br />

fracturing activities, or other sources of contamination.<br />

7.5.4. Status and Preliminary Data<br />

Two rounds of sampling have been completed: the first in July 2011 and the second in March 2012.<br />

During July 2011, 13 domestic wells and three surface water locations (small streams and spring<br />

discharges) were sampled. During March 2012, 13 domestic wells and two surface water locations<br />

were sampled. The locations of sampling sites are displayed in Figure 35.<br />

7.5.5. Next Steps<br />

Additional sampling rounds will be conducted to verify data collected from the first two rounds of<br />

sampling. Additional sampling locations may be included in the future as analytical data is<br />

evaluated and additional pertinent information becomes available. More focused investigations<br />

may also be conducted, if warranted, at locations where impacts associated with hydraulic<br />

fracturing may have occurred.<br />

7.5.6. Quality Assurance Summary<br />

The initial QAPP for this case study, “Hydraulic Fracturing Retrospective Case Study, Marcellus<br />

Shale, Washington County, PA,” was approved by the designated EPA QA Manager on July 21, 2011<br />

(US EPA, 2012n). A revision to the QAPP was made before the second sampling event and was<br />

151


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

approved on March 5, 2012, to update project organization, lab accreditation information, sampling<br />

methodology, to add radium isotope analyses and gross alpha/beta analyses, to modify critical<br />

analytes, and to change the analytical method for determining water isotope values. There have<br />

been no significant deviations from the QAPP during any sampling event, and therefore no impact<br />

on data quality. A field TSA was conducted on March 26, 2011; no findings were identified. See<br />

Section 7.1.1 for information related to the laboratory TSAs.<br />

As results are reported and raw data are provided from the laboratories, ADQs are performed to<br />

verify that the quality requirements specified in the approved QAPP were met. Data will be<br />

qualified if necessary, based on these ADQs. The results of these ADQs will be reported in the final<br />

report on this project.<br />

152


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

7.6. Wise County, Texas<br />

7.6.1. Project Introduction<br />

Wise County, Texas, is mostly rural, with a total population of about 60,000 and about 66 people<br />

per square mile (USCB, 2010f). Current gas development activities in Wise County are in the<br />

Barnett Shale, which is an unconventional shale in the Fort Worth Basin adjoining the Bend Arch<br />

Basin of north-central Texas. Figure 36 shows the extent of the Barnett Shale in Texas. In recent<br />

years, gas production in Wise County has increased due to improvements in horizontal drilling and<br />

hydraulic fracturing technologies.<br />

Figure 36. Extent of the Barnett Shale in north-central Texas (US EIA, 2011e; USCB, 2012a, c). The case study<br />

focuses on three distinct locations within Wise County.<br />

The intent of this case study is to investigate homeowner concerns about changes in the ground<br />

water quality in Wise County that may be related to the recent increase in the hydraulic fracturing<br />

of oil and gas wells. Sampling locations in Wise County were chosen based on reported complaints<br />

of changes in drinking water quality and are clustered in three distinct locations: two near Decatur<br />

and one near Alvord. Homeowners in the two locations near Decatur reported changes in water<br />

153


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

quality, including changes in turbidity, color, smell, and taste. Homeowners near Alvord also<br />

reported changes in drinking water quality, although no more specific concerns were identified.<br />

Concerns about potential hydraulic fracturing impacts to ground water resources in Wise County<br />

are related to flowback fluid discharge to shallow aquifers, gas migration to shallow aquifers, spills<br />

on well pads, and leaking impoundments. Residential or agricultural practices, or aquifer<br />

drawdown unrelated to oil and gas development, may also be sources of ground water<br />

contamination at these sites.<br />

7.6.2. Site Background<br />

Geology. Wise County is located in the Bend Arch-Fort Worth Basin, which was formed during the<br />

late Paleozoic Ouachita Orogeny by the convergence of Laurussia and Gondwana in a narrow,<br />

restricted, inland seaway (Bruner and Smosna, 2011). The stratigraphy of the Bend Arch-Fort<br />

Worth Basin is characterized by limestones, sandstones, and shales. The Barnett Shale is of<br />

Mississippian age (Appendix D) and extends throughout the Bend Arch-Fort Worth Basin: south<br />

from the Muenster Arch, near the Oklahoma border, to the Llano Uplift in Burnet County and west<br />

from the Ouachita Thrust Front, near Dallas, to Taylor County (Bruner and Smosna, 2011). The<br />

Barnett Shale ranges from about 50 to 1,000 feet thick and occurs at depths ranging from 4,000 to<br />

8,500 feet (Bruner and Smosna, 2011). In the northeastern portion of the Fort Worth Basin, the<br />

Barnett Shale is divided by the presence of the Forestburg Limestone, but this formation tapers out<br />

toward the southern edge of Wise County (Bruner and Smosna, 2011). The Barnett Shale is<br />

bounded by the Chappel Limestone below it and the Marble Falls Limestone above it (Bruner and<br />

Smosna, 2011). A recent estimate of the potential total gas yield was 820 billion cubic feet of gas per<br />

square mile, which is a significant increase from earlier estimates (Bruner and Smosna, 2011).<br />

Water Resources. Wise County is drained by the Trinity River. Residents in the county often rely on<br />

the Trinity Aquifer as a major source of drinking water. In addition to drinking water, the Trinity<br />

Aquifer is also used for irrigation, industrial water, and hydraulic fracturing source water. The<br />

aquifer is composed of three formations, deposited in the Cretaceous: Paluxy, Glen Rose, and Twin<br />

Mountain (Nordstrom, 1982; Reutter and Dunn, 2000; Scott and Armstrong, 1932). In the northern<br />

part of Wise County, the Glen Rose formation pinches out, leaving only the Paluxy and Twin<br />

Mountain Formations, which together are occasionally referred to as the Antlers Formation<br />

(Nordstrom, 1982; Reutter and Dunn, 2000). The composition of the Paluxy Formation is fine sand,<br />

sandy shale, and shale and yields small to moderate quantities of water (Nordstrom, 1982). The<br />

Glen Rose Formation is composed of limestone, marl, shale, and anhydrite. The Glen Rose yields<br />

small quantities of water in localized areas (Nordstrom, 1982). Finally, the composition of the Twin<br />

Mountain Formation is fine to coarse sand, shale, clay, and basal gravel and conglomerate. This<br />

formation yields moderate to large quantities of water (Nordstrom, 1982). The Trinity Aquifer is<br />

overlain by the Walnut Creek Formation and is underlain by Graham Formation, both of which act<br />

as confining layers (Scott and Armstrong, 1932). Before modern water usage, it was artesian.<br />

Table 54 summarizes background water quality data for the Trinity Aquifer in Wise County<br />

(Reutter and Dunn, 2000). The water quality is expected to be slightly different in the northern<br />

portion of the county than the southern portion of the county due to the “pinching out” of the Glen<br />

Rose Formation. From the reported data, the major water types in Wise County are calcium<br />

154


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

bicarbonate, calcium chloride, and sodium bicarbonate (Reutter and Dunn, 2000). All three water<br />

types are present in northern Wise County, but only the calcium bicarbonate and calcium chloride<br />

water types were observed in southern Wise County. The data collected at study locations will be<br />

compared to this compiled background data as part of the initial screening to determine if any<br />

contamination has occurred in study locations.<br />

Table 54. Background water quality data for all of Wise County, Texas, and its northern and southern regions<br />

(Reutter and Dunn, 2000). Range of concentrations shown, with median values reported in parentheses.<br />

Parameter<br />

Units<br />

Wise County<br />

Concentration Ranges<br />

North Wise<br />

County<br />

South Wise<br />

County<br />

Alkalinity mg CaCO 3 /L 130–430 (335) 190–430 (330) 130–420 (360)<br />

Aluminum µg/L 1–5 (2) 2–5 (2) 1–5 (2)<br />

Ammonia mg N/L


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

Parameter<br />

Specific<br />

conductance<br />

Units<br />

Wise County<br />

Concentration Ranges<br />

North Wise<br />

County<br />

South Wise<br />

County<br />

µS/cm 710–4,590 (913) 71–4,590 (911) 510–2,380 (914)<br />

Sulfate mg/L 10–250 (46) 26–250 (45) 10–160 (46)<br />

Uranium µg/L


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Figure 37. Location of sampling sites in Wise County, Texas.<br />

Water samples collected at these locations are being analyzed for the chemicals listed in Section<br />

7.1.1 as well as the chemicals listed in the QAPP (US EPA, 2012p). Together these measurements<br />

support the objective of determining if ground water resources have been impacted by hydraulic<br />

fracturing activities, or other sources of contamination.<br />

7.6.4. Status and Preliminary Data<br />

Two rounds of sampling have been conducted at all locations in Wise County: one round in<br />

September 2011 and one round in March 2012. The September 2011 sampling event included 11<br />

domestic wells, one industrial well, and three surface water (pond) samples. The March 2012<br />

sampling event included the same wells as the September 2011 sampling event, with an additional<br />

four domestic wells and the loss of one domestic well. The locations of all sampling sites are<br />

displayed in Figure 37.<br />

7.6.5. Next Steps<br />

Additional sampling rounds will be conducted to verify data collected from the first two rounds of<br />

sampling. Additional sampling locations may be included in the future as analytical data are<br />

evaluated and additional pertinent information becomes available. More focused investigations<br />

may also be conducted, if warranted, at locations where impacts may have occurred.<br />

157


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

7.6.6. Quality Assurance Summary<br />

The initial QAPP for this case study, “Hydraulic Fracturing Retrospective Case Study, Wise, TX,” was<br />

approved by the designated EPA QA Manager on June 20, 2011 (US EPA, 2012p). A revision to the<br />

QAPP was made before the second sampling event and was approved on February 27, 2012. The<br />

revision included the addition of isotopic analysis, USGS laboratory information, 82 revised sampling<br />

locations, Region 8 laboratory accreditation status, geophysical measurement methods and QC, data<br />

qualifiers, personnel changes, and analytical method updates. A second revision was approved on<br />

May 25, 2012, for the next sampling event to include the Phase 2 sampling information, the method<br />

for qualifying field blanks, and the modified sampling schedule. The second QAPP revision also<br />

replaced EPA Method 200.7 with 6010C and replaced metals QC criteria with revised criteria. A<br />

third revision to the QAPP was approved on September 10, 2012, to add information on March<br />

2012 sampling, add strontium and stable water isotopes to analytes list, and delete diesel range<br />

organics and gasoline range organics. The third QAPP revision also replaced EPA Method 6010C<br />

with 200.7. 83 There have been no significant deviations from the QAPP during any sampling event,<br />

and therefore no impact on data quality. A field TSA was conducted on September 21, 2011; no<br />

findings were identified. See Section 7.1.1 for information related to the laboratory TSAs.<br />

As results are reported and raw data are provided from the laboratories, ADQs are performed to<br />

verify that the quality requirements specified in the approved QAPP were met. Data will be<br />

qualified if necessary, based on these ADQs. The results of these ADQs will be reported in the final<br />

report on this project.<br />

82 USGS provided isotope support for the Wise County retrospective case study. A detailed account of the role of USGS can<br />

be found in Appendix A of the Wise County QAPP.<br />

83<br />

EPA Method 200.7 was referenced in the initial QAPP and the first QAPP revision. It was changed in the second QAPP<br />

revision to EPA Method 6010C, but since then it was determined by QA staff that the use of 200.7 as the “base” method<br />

was appropriate as 200.7 incorporates 6010C by reference.<br />

158


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

8. Conducting High-Quality Science<br />

The EPA ensures that its research activities result in high-quality science through the use of QA and<br />

peer review activities. Specific QA activities performed by the EPA ensure that the agency’s<br />

environmental data are of sufficient quantity and quality to support the data’s intended use. Peer<br />

review ensures that the data are sound and used appropriately. The use of QA measures and peer<br />

review helps ensure that the EPA conducts high-quality science that can be used to inform<br />

policymakers, industry, and the public.<br />

8.1. Quality Assurance<br />

All agency research projects that generate or use environmental data to make conclusions or<br />

recommendations must comply with the EPA QA program requirements. The EPA laboratories and<br />

external organizations involved with the generation or use of environmental data are supported by<br />

QA professionals who oversee the implementation of the QA program for their organization. To<br />

ensure scientifically defensible results, this study complies with the agency-wide Quality Policy CIO<br />

2106 (US EPA, 2008), EPA Order CIO 2105.0 (US EPA, 2000a, c), the EPA’s Information Quality<br />

Guidelines (US EPA, 2002), the EPA’s Laboratory Competency Policy (US EPA, 2004a), and Chapter<br />

13 of the Office of Research and Development’s Policies and Procedures Manual (US EPA, 2006).<br />

Given the cross-organizational nature of this study, a Quality Management Plan was developed (US<br />

EPA, 2012t) and a Program QA Manager was chosen to coordinate a rigorous QA approach and<br />

oversee its implementation across all participating organizations within the EPA. The Quality<br />

Management Plan defines the QA-related policies, procedures, roles, responsibilities, and<br />

authorities for the study and documents how the EPA will plan, implement, and assess the<br />

effectiveness of its QA and QC operations. In light of the importance and organizational complexity<br />

of the study, the Quality Management Plan was created to make certain that all research be<br />

conducted with integrity and strict quality controls.<br />

The Quality Management Plan sets forth the following rigorous QA approach:<br />

• Individual research projects must comply with agency requirements and guidance for<br />

QAPPs.<br />

• TSAs and audits of data quality will be conducted for individual research projects as<br />

described in the QAPPs.<br />

• Performance evaluations of analytical systems will be conducted.<br />

• Products will undergo QA review. Applicable products may include reports, journal<br />

articles, symposium/conference papers, extended abstracts, computer products/<br />

software/models/databases, and scientific data.<br />

• Reports will have readily identifiable QA sections.<br />

Research records will be managed according to EPA Records Schedule 501, “Applied and Directed<br />

Scientific Research”(US EPA, 2011c).<br />

159


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

The Quality Management Plan applies to all research activities conducted under the EPA’s Study of<br />

the Potential Impacts of Hydraulic Fracturing on Drinking Water Resources. More information about<br />

specific QA protocols, including management, organization, quality-system components, personnel<br />

qualification and training, procurement of items and services, documentation and records,<br />

computer requirements, planning, implementation, assessment, and quality improvement, can be<br />

found in the Quality Management Plan. 84<br />

Project-specific details of individual research projects are documented in a QAPP. All work<br />

performed or funded by the EPA that involves the acquisition of environmental data must have an<br />

approved QAPP. The QAPP documents the planning, implementation, and assessment procedures<br />

for a particular project, as well as any specific QA and QC activities. It integrates all the technical<br />

and quality aspects of the project in order to provide a guide for obtaining the type and quality of<br />

environmental data and information needed for a specific decision or use. Quality assurance project<br />

plans are living documents that undergo revisions as needed. Individual QAPPs for the various<br />

research projects included in this study are available on the study website<br />

(http://www.epa.gov/<strong>hf</strong>study) and are summarized in Appendix C.<br />

Regular technical assessments of project operation, systems, and data related to the study are<br />

conducted as detailed in the Quality Management Plan. A technical assessment is “a systematic and<br />

objective examination of a project to determine whether environmental data collection activities<br />

and related results comply with the project’s QAPP, whether the activities are implemented<br />

effectively, and whether they are sufficient and adequate to achieve QAPP’s data quality goals” (US<br />

EPA, 2000b). Assessment components include quality system assessments, technical system<br />

assessments, verification of data, audits of data quality, and surveillance. More details about<br />

assessments and audits required for this study can be found in the Quality Management Plan and<br />

project-specific QAPPs.<br />

Quality Assurance and Projects Involving the Generation of New Data. Research projects that<br />

generate new data (e.g., case studies, laboratory studies, some toxicity assessments) will contribute<br />

to the growing body of scientific literature about environmental issues associated with hydraulic<br />

fracturing. The QA/QC procedures detailed in these QAPPs meet the requirements of the hydraulic<br />

fracturing Quality Management Plan, detailed above, and also focus on those practices necessary for<br />

assuring the quality of measurement data generated by the EPA. Samples must be collected,<br />

preserved, transported, and stored in a manner that retains their integrity; these issues are<br />

addressed in individual QAPPs. Also described in QAPPs are the methods used for sample analysis,<br />

including details about the appropriate frequency of calibration of analytical instrumentation and<br />

measurement devices. Quality control samples are identified that can be used to check for potential<br />

contamination of samples and to check for measurement errors that can be caused by difficult<br />

sample matrices. The QAPPs for generation of new data provide details on the logistics of who,<br />

where, when and how new data will be generated.<br />

84 Research initiated prior to the implementation of the study-specific Quality Management Plan was conducted under<br />

Quality Management Plans associated with each of the EPA Office of Research and Development’s individual labs and<br />

centers.<br />

160


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Quality Assurance and Projects Involving Existing Data. Research projects that involve acquiring and<br />

analyzing existing data (i.e., data that are not new data generated by or for the EPA) must conform<br />

to the requirements of the Quality Management Plan, including the development of a QAPP. The<br />

focus of QAPPs for existing data is on setting criteria that will filter out any data that are of<br />

insufficient quality to meet project needs. This starts with describing the process for locating and<br />

acquiring the data. How the data will be evaluated for their planned use and how the integrity of the<br />

data will be maintained throughout the collection, storing, evaluation, and analysis processes are<br />

also important features of a QAPP for existing data.<br />

Quality Assurance and Report Preparation. Quality assurance requirements also extend to the two<br />

primary products of this study: this progress report and the report of results. As required by the<br />

Quality Management Plan, this progress report has undergone QA review before its release, and the<br />

report of results will do the same. These requirements serve to ensure that the reports are<br />

defensible and scientifically sound.<br />

8.2. Peer Review<br />

Peer review, an important part of every scientific study, is a documented critical review of a specific<br />

scientific and/or technical work product (e.g., paper, report, presentation). It is an in-depth<br />

assessment of the assumptions, calculations, extrapolations, alternate interpretations,<br />

methodology, acceptance criteria, and conclusions in the work product and the documents that<br />

support them. Peer review is conducted by individuals (or organizations) independent of those who<br />

performed the work and equivalent in technical expertise (US EPA, 2012e; US OMB, 2004).<br />

Feedback from the review process is used to revise the draft product to make certain the final work<br />

product reflects sound technical information and analyses.<br />

Peer review can take many forms depending on the nature of the work product. Work products<br />

generated through the EPA’s Study of the Potential Impacts of Hydraulic Fracturing on Drinking<br />

Water Resources will be subjected to both internal and external peer review. Internal peer review<br />

occurs when work products are reviewed by independent experts within the EPA, while external<br />

peer review engages experts outside of the agency, often through scientific journals, letter reviews,<br />

or ad hoc panels.<br />

The EPA often engages the Science Advisory Board, an external federal advisory committee, to<br />

conduct peer reviews of high-profile scientific matters relevant to the agency. Members of an ad hoc<br />

panel convened under the auspices of the Science Advisory will provide comment on this progress<br />

report. 85 Panel members are nominated by the public and chosen based on factors such as technical<br />

expertise, knowledge, experience, and absence of any real or perceived conflicts of interest to<br />

create a balanced review panel. In August 2012, the EPA issued a Federal Register notice requesting<br />

public nominations for technical experts to form a Science Advisory Board ad hoc panel to provide<br />

advice on the status of the research described in this progress report (US EPA, 2012v). This panel is<br />

85 Information about this process is available at http://yosemite.epa.gov/sab/sabproduct.nsf/<br />

02ad90b136fc21ef85256eba00436459/b436304ba804e3f885257a5b00521b3b!OpenDocument.<br />

161


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

also expected to review the report of results, which has been classified as a Highly Influential<br />

Scientific Assessment. 86<br />

86 The Office of Management and Budget’s Peer Review Bulletin (US OMB, 2004) defines Highly Influential Scientific<br />

Assessments as scientific assessments that could (1) have a potential impact of more than $500 million in any year or (2)<br />

are novel, controversial, or precedent-setting or have significant interagency interest. The Peer Review Bulletin describes<br />

specific peer review requirements for Highly Influential Scientific Assessments.<br />

162


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

9. Research Progress Summary and<br />

Next Steps<br />

This report describes the progress made for each of the research projects conducted as part of the<br />

EPA’s Study of the Potential Impacts of Hydraulic Fracturing on Drinking Water Resources. This<br />

chapter provides an overview of the progress made for each research activity as well as the<br />

progress made for each stage of the water cycle presented in Section 2.1. It also describes, in more<br />

detail, the report of results.<br />

9.1. Summary of Progress by Research Activity<br />

The EPA is using a transdisciplinary research approach to investigate the potential relationship<br />

between hydraulic fracturing and drinking water resources. This approach includes compiling and<br />

analyzing data from existing sources, evaluating scenarios using computer models, carrying out<br />

laboratory studies, assessing the toxicity associated with hydraulic fracturing-related chemicals,<br />

and conducting case studies.<br />

Analysis of Existing Data. To date, data from seven sources have been obtained for review and<br />

ongoing analysis, including:<br />

• Information provided by nine hydraulic fracturing service companies.<br />

• 333 well files supplied by nine oil and gas operators.<br />

• Over 12,000 chemical disclosure records from FracFocus, the national hydraulic fracturing<br />

chemical registry managed by the Ground Water Protection Council and the Interstate Oil<br />

and Gas Compact Commission.<br />

• Spill reports from four different sources, including databases from the National Response<br />

Center, Colorado, New Mexico, and Pennsylvania.<br />

As part of its literature review, the EPA has compiled, and continues to search for, literature<br />

relevant to the secondary research questions listed in Section 2.1. This includes documents<br />

provided by stakeholders and recommended by the Science Advisory Board during its review of the<br />

draft study plan. 87 A Federal Register notice requesting peer-reviewed data and publications<br />

relevant to the study, including information on advances in industry practices and technologies, has<br />

recently been published (US EPA, 2012u).<br />

Scenario Evaluations. Potential impacts to drinking water sources from withdrawing large volumes<br />

of water in both a semi-arid and a humid river basin—the Upper Colorado River Basin in the west<br />

and the Susquehanna River Basin in the east—are being assessed. Additionally, complex computer<br />

models are being used to explore the possibility of subsurface gas and fluid migration from deep<br />

shale formations to overlying aquifers in six different scenarios. These scenarios include poor well<br />

87 Additional information on the Science Advisory Board review of the EPA’s Draft Plan to Study the Potential Impacts of<br />

Hydraulic Fracturing on Drinking Water Resources is available at http://www.epa.gov/<strong>hf</strong>study/peer-review.html.<br />

163


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

construction and hydraulic communication via fractures (natural and created) and nearby existing<br />

wells. As a first step, the subsurface migration simulations will examine realistic scenarios to assess<br />

the conditions necessary for hydraulic communication rather than the probability of migration<br />

occurring. In a separate research project, the EPA is using surface water transport models to<br />

estimate concentrations of bromide and radium at public water supply intakes downstream from<br />

wastewater treatment facilities that discharge treated hydraulic fracturing wastewater.<br />

Laboratory Studies. The ability to analyze and determine the presence and concentration of<br />

chemicals in environmental samples is critical to the EPA’s study. In most cases, standard EPA<br />

methods are being used for laboratory analyses. In other cases, however, standard methods do not<br />

exist for the low-level detection of chemicals of interest or for use in the complex matrices<br />

associated with hydraulic fracturing wastewater. Where necessary, existing analytical methods are<br />

being tested, modified, and verified for use in this study and by others. Analytical methods are<br />

currently being tested and modified for several classes of chemicals, including glycols, acrylamides,<br />

ethoxylated alcohols, DBPs, radionuclides, and inorganic chemicals.<br />

Laboratory studies focusing on the potential impacts of inadequate treatment of hydraulic<br />

fracturing wastewater on drinking water resources are being planned and conducted. The studies<br />

include assessing the ability of hydraulic fracturing wastewater to create brominated DBPs and<br />

testing the efficacy of common wastewater treatment processes on removing selected<br />

contaminants from hydraulic fracturing wastewater. Samples of surface water, raw hydraulic<br />

fracturing wastewater, and treated effluent have been collected for the source apportionment<br />

studies, which aim to identify the source of high chloride and bromide levels in rivers accepting<br />

treated hydraulic fracturing wastewater.<br />

Toxicity Assessment. The EPA has evaluated data to identify chemicals reportedly used in hydraulic<br />

fracturing fluids from 2005 to 2011 and chemicals found in flowback and produced water.<br />

Appendix A contains tables of these chemicals, with over 1,000 chemicals identified. Chemical,<br />

physical, and toxicological properties have been compiled for chemicals with known chemical<br />

structures. Existing models are being used to estimate properties in cases where information is<br />

lacking. At this time, the EPA has not made any judgment about the extent of exposure to these<br />

chemicals when used in hydraulic fracturing fluids or found in hydraulic fracturing wastewater, or<br />

their potential impacts on drinking water resources.<br />

Case Studies. Two rounds of sampling at all five retrospective case study locations have been<br />

completed. In total, water samples have been collected from over 70 domestic water wells, 15<br />

monitoring wells, and 13 surface water sources, among others. A third round of sampling is<br />

expected to occur this fall in Las Animas and Huerfano Counties, Colorado; Dunn County, North<br />

Dakota; and Wise County, Texas. Additional sampling in Bradford and Washington Counties,<br />

Pennsylvania, is projected to take place in spring 2013.<br />

The EPA continues to work with industry partners to plan and begin research activities for<br />

prospective case studies.<br />

164


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

9.2. Summary of Progress by Water Cycle Stage<br />

Figures 38 and 39 illustrate the research underway for each stage of the hydraulic fracturing water<br />

cycle. The fundamental research questions and research focus areas are briefly described below for<br />

each water cycle stage; for more detail on the stages of the hydraulic fracturing water cycle and<br />

their associated research projects, see Section 2.1.<br />

Water Acquisition: What are the possible impacts of large volume water withdrawals from ground<br />

and surface waters on drinking water resources Work in this area focuses on understanding the<br />

volumes and sources of water needed for hydraulic fracturing operations, and the potential impacts<br />

of water withdrawals on drinking water quantity and quality. Effects of recently emerging trends in<br />

water recycling will be considered in the report of results.<br />

Chemical Mixing: What are the possible impacts of surface spills on or near well pads of hydraulic<br />

fracturing fluids on drinking water resources Spill reports from several databases are being<br />

reviewed to identify volumes and causes of spills of hydraulic fracturing fluids and wastewater.<br />

Information on the chemicals used in hydraulic fracturing fluids and their known chemical,<br />

physical, and toxicological properties has been compiled.<br />

Well Injection: What are the possible impacts of the injection and fracturing process on drinking water<br />

resources Work currently underway is focused on identifying conditions that may be associated<br />

with the subsurface migration of gases and fluids to drinking water resources. The EPA is exploring<br />

gas and fluid migration due to inadequate well construction as well as the presence of nearby<br />

natural faults and fractures or man-made wells.<br />

Flowback and Produced Water: What are the possible impacts of surface spills on or near well pads of<br />

flowback and produced water on drinking water resources As with chemical mixing, research in this<br />

area focuses on reviewing spill reports of flowback and produced water as well as collecting<br />

information on the composition of hydraulic fracturing wastewater. Known chemical, physical, and<br />

toxicological properties of the components of flowback and produced water are being compiled.<br />

Wastewater Treatment and Waste Disposal: What are the possible impacts of inadequate treatment of<br />

hydraulic fracturing wastewater on drinking water resources Work in this area focuses on<br />

evaluating treatment and disposal practices for hydraulic fracturing wastewater. Since some<br />

wastewater is known to be discharged to surface water after treatment in POTWs or commercial<br />

treatment systems, the EPA is investigating the efficacy of common treatment processes at<br />

removing selected components in flowback and produced water. Potential impacts to downstream<br />

public water supplies from discharge of treated hydraulic fracturing wastewater are also being<br />

investigated, including the potential for the formation of Br-DBPs.<br />

165


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Water Acquisition Chemical Mixing Well Injection<br />

Literature Review<br />

Review and summarize literature on:<br />

• Volumes and sources of water used in<br />

hydraulic fracturing fluids<br />

• Local impacts to water availability in<br />

areas with hydraulic fracturing activity<br />

• Water quality impacts from ground<br />

and surface water withdrawals<br />

Service Company Analysis<br />

Summarize data provided by nine<br />

hydraulic fracturing service companies<br />

on volumes and sources of water used<br />

in hydraulic fracturing fluids<br />

Well File Review<br />

Summarize data from 333 well files on<br />

volumes and source of water used in<br />

hydraulic fracturing fluids<br />

FracFocus Analysis<br />

Compile and summarize total water<br />

volumes reported in FracFocus by<br />

geographic location, well depth, water<br />

types, and oil/gas production<br />

Literature Review<br />

Review and summarize literature on:<br />

• Spills of hydraulic fracturing fluids or<br />

chemical additives<br />

• Chemicals used in hydraulic fracturing<br />

fluids<br />

• Environmental fate and transport of<br />

selected chemicals in hydraulic<br />

fracturing fluids<br />

Spills Database Analysis<br />

Compile and evaluate spill information<br />

from three state databases (CO, NM,<br />

PA) and one national database (NRC)<br />

Service Company Analysis<br />

Evaluate information on:<br />

• Spills of hydraulic fracturing fluids or<br />

chemical additives<br />

• Chemicals used in hydraulic fracturing<br />

fluids from 2005 to 2010<br />

Well File Review<br />

Evaluate spill data from 333 well files<br />

FracFocus Analysis<br />

Compile a list of chemicals reported in<br />

FracFocus and summarize chemical<br />

usage by frequency and geographic<br />

location<br />

Literature Review<br />

Review and summarize literature on<br />

possible subsurface migration due to:<br />

• Faulty well construction<br />

• Nearby natural or man-made conduits<br />

Service Company Analysis<br />

Review and summarize standard<br />

operating procedures for information on:<br />

• Practices related to establishing the<br />

mechanical integrity of wells being<br />

hydraulically fractured<br />

• Procedures used during injection of<br />

the fracturing fluid<br />

Well File Review<br />

Review well construction data found in<br />

well files to assess the effectiveness of<br />

current well construction practices at<br />

isolating the wellbore from surrounding<br />

ground water<br />

Analysis of Existing Data<br />

Scenario Evaluations<br />

Laboratory Studies<br />

Toxicity Assessment<br />

Case Studies<br />

Figure 38a. Summary of research projects underway for the first three stages of the hydraulic fracturing water cycle.<br />

166


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Water Acquisition Chemical Mixing Well Injection<br />

Water Availability Modeling<br />

• Summarize data on water usage for<br />

hydraulic fracturing in a semi-arid<br />

climate (Upper Colorado River Basin)<br />

and a humid climate (Susquehanna<br />

River Basin)<br />

• Use watershed models to explore<br />

water availability for public water<br />

supplies under a variety of scenarios,<br />

focusing on water usage in the Upper<br />

Colorado and Susquehanna River<br />

Basins<br />

Analysis of Existing Data<br />

Scenario Evaluations<br />

Laboratory Studies<br />

Toxicity Assessment<br />

Case Studies<br />

Analytical Method Development<br />

Develop analytical methods for the<br />

detection of selected chemicals reported<br />

to be in hydraulic fracturing fluids<br />

Toxicity Assessment<br />

Compile or estimate chemical, physical,<br />

and toxicological properties for<br />

chemicals with known chemical<br />

structures that are reported to be in<br />

hydraulic fracturing fluids<br />

Retrospective Case Studies<br />

Consider spills of hydraulic fracturing<br />

fluids as a possible source of reported<br />

changes in water quality of local<br />

drinking water wells<br />

Subsurface Migration Modeling<br />

Apply computer models to explore the<br />

potential for gas or fluid migration from:<br />

• Incomplete well cementing or cement<br />

failure during hydraulic fracturing<br />

• Nearby wells and existing faults<br />

Retrospective Case Studies<br />

Consider potential impacts to drinking<br />

water sources from:<br />

• Relatively shallow hydraulic fracturing<br />

operations<br />

• Release of hydraulic fracturing fluids<br />

during the injection process<br />

• Poor well construction practices<br />

Figure 38b. Summary of research projects underway for the first three stages of the hydraulic fracturing water cycle.<br />

167


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Flowback and Produced Water<br />

Literature Review<br />

Review and summarize literature on:<br />

• Spills of flowback and produced water<br />

• Chemicals found in hydraulic<br />

fracturing wastewater<br />

• Environmental fate and transport of<br />

selected chemicals in hydraulic<br />

fracturing wastewater<br />

Spills Database Analysis<br />

Compile spill information from three<br />

state databases (CO, NM, PA) and one<br />

national database (NRC)<br />

Service Company Analysis<br />

Evaluate information on:<br />

• Spills of flowback and produced water<br />

• Chemicals detected in hydraulic<br />

fracturing wastewater<br />

Well File Review<br />

Evaluate spill data from 333 well files<br />

Wastewater Treatment<br />

and Waste Disposal<br />

Literature Review<br />

Review and summarize literature on:<br />

• Disposal practices associated with<br />

hydraulic fracturing wastewater<br />

• The treatability of hydraulic fracturing<br />

wastewater<br />

• Potential impacts to drinking water<br />

treatment facilities from surface<br />

discharge of treated hydraulic<br />

fracturing wastewater<br />

Well File Review<br />

Summarize data from 333 well files on<br />

the volume and final disposition of<br />

flowback and produced water<br />

FracFocus Analysis<br />

Summarize data on water types<br />

reported in FracFocus by volume and<br />

geographic location, focusing on<br />

recycled water<br />

Analysis of Existing Data<br />

Scenario Evaluations<br />

Laboratory Studies<br />

Toxicity Assessment<br />

Case Studies<br />

Figure 39a. Summary of research projects underway for the last two stages of the hydraulic fracturing water cycle.<br />

168


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Flowback and Produced Water<br />

Analytical Method Development<br />

Develop analytical methods for the<br />

detection of selected chemicals in<br />

hydraulic fracturing wastewater matrices<br />

Toxicity Assessment<br />

Compile or estimate chemical, physical,<br />

and toxicological properties for<br />

chemicals reported to be in hydraulic<br />

fracturing wastewater<br />

Retrospective Case Studies<br />

Consider spills or leaks of hydraulic<br />

fracturing wastewater as a possible<br />

source of reported changes in water<br />

quality of local drinking water wells<br />

Wastewater Treatment<br />

and Waste Disposal<br />

Surface Water Modeling<br />

Apply computer models to calculate<br />

downstream concentrations of selected<br />

contaminants at public water intakes<br />

under a variety of scenarios<br />

Source Apportionment Studies<br />

Collect samples from two wastewater<br />

treatment facilities and river networks<br />

and use computer models to identify the<br />

contribution of hydraulic fracturing<br />

wastewater to chemical concentrations<br />

found at downstream public water<br />

intakes<br />

Wastewater Treatability Studies<br />

Conduct laboratory experiments to<br />

identify the fate of selected chemicals<br />

found in flowback in common treatment<br />

processes, including conventional,<br />

commercial and water reuse processes<br />

Analysis of Existing Data<br />

Scenario Evaluations<br />

Laboratory Studies<br />

Toxicity Assessment<br />

Case Studies<br />

Br-DBP Precursor Studies<br />

Conduct laboratory studies on the<br />

potential for treated hydraulic fracturing<br />

wastewater to form Br-DBPs during<br />

common drinking water treatment<br />

processes<br />

Figure 39b. Summary of research projects underway for the last two stages of the hydraulic fracturing water cycle.<br />

169


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

9.3. Report of Results<br />

This is a status report, describing the current progress made on the research projects that make up<br />

the agency’s Study of the Potential Impacts of Hydraulic Fracturing on Drinking Water Resources.<br />

Results from individual research projects will undergo peer review prior to publication either as<br />

articles in scientific journals or EPA reports. The EPA plans to synthesize results from the published<br />

reports with a critical literature review in a report of results that will answer as completely as<br />

possible the research questions identified in the Study Plan. The report of results has been<br />

determined to be a Highly Influential Scientific Assessment and will undergo peer review by the<br />

Science Advisory Board. Ultimately, the results of this study are expected to inform the public and<br />

provide policymakers at all levels with high-quality scientific knowledge that can be used in<br />

decision-making processes.<br />

The report of results will also be informed by information provided through the ongoing<br />

stakeholder engagement process described in Section 1.1. This process is anticipated to provide<br />

agency scientists with updates on changes in industry practices and technologies relevant to the<br />

study. While the EPA expects hydraulic fracturing technology to develop between now and the<br />

publication of the report of results, the agency believes that the research described here will<br />

provide timely information that will contribute to the state of knowledge on the relationship<br />

between hydraulic fracturing and drinking water resources. For example, some companies may<br />

adopt new injection or wastewater treatment technologies and practices, while others may<br />

continue to use current technologies and practices. Many of the practices, including wastewater<br />

treatment and disposal technologies used by POTWs, are not expected to change significantly<br />

between now and the report of results.<br />

Results from the study are expected to identify potential impacts to drinking water resources, if<br />

any, from water withdrawals, the fate and transport of chemicals associated with hydraulic<br />

fracturing, and wastewater treatment and waste disposal. Information on the toxicity of hydraulic<br />

fracturing-related chemicals is also being gathered. Although these data may be used to assess the<br />

potential risks to drinking water resources from hydraulic fracturing activities, the report of results<br />

is not intended to quantify risks. Results presented in the report of results will be appropriately<br />

discussed and all uncertainties will be described.<br />

The EPA will strive to make the report of results as clear and definitive as possible in answering all<br />

of the primary and secondary research questions, at that time. Science and technology evolve,<br />

however: the agency does not believe that the report of results will provide definitive answers on<br />

all research questions for all time and fully expects that additional research needs will be identified.<br />

9.4. Conclusions<br />

This report presents the EPA’s progress in conducting its Study of the Potential Impacts of Hydraulic<br />

Fracturing on Drinking Water Resources. Chapters 3 through 7 provide individual progress reports<br />

for each of the research projects that make up this study. Each project progress report describes the<br />

project’s relationship to the study, research methods, and status and summarizes QA activities.<br />

Information presented as part of this report cannot be used to draw conclusions about potential<br />

impacts to drinking water resources from hydraulic fracturing.<br />

170


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

The EPA is committed to conducting a study that uses the best available science, independent<br />

sources of information, and a transparent, peer-reviewed process that ensures the validity and<br />

accuracy of the results. The EPA will seek input from individual members of an ad hoc expert panel<br />

convened under the auspices of the EPA’s Science Advisory Board. Information about this process is<br />

available at http://yosemite.epa.gov/sab/sabproduct.nsf/02ad90b136fc21ef85256eba00436459/<br />

b436304ba804e3f885257a5b00521b3b!OpenDocument. The individual members of the ad hoc<br />

panel will consider public comment. The EPA will consider feedback from the individual experts, as<br />

informed by the public’s comments, in the development of the report of results.<br />

171


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

10. References<br />

Abbott, P. O. 1985. Description of Water-Systems Operations in the Arkansas River Basin, Colorado.<br />

Water-Resources Investigations Report 85-4092. US Geological Survey. 79 p. Available at<br />

http://pubs.er.usgs.gov/publication/wri854092. Accessed November 14, 2012.<br />

Abbott, P. O., Geldon, A. L., Cain, D., Hall, A. P. and Edelmann, P. 1983. Hydrology of Area 61,<br />

Northern Great Plains and Rocky Mountain Coal Provinces, Colorado and New Mexico.<br />

Open-File Report 83-132. US Geological Survey. 105 p. Available at http://pubs.er.usgs.gov/<br />

publication/ofr83132. Accessed November 14, 2012.<br />

Toxicity Prediction by Komputer Assisted Technology (TOPKAT) version 3.5. 2012. Accelrys<br />

Software, Inc., San Diego, California.<br />

ChemFolder version 12.00. 2008. ACD Labs, Toronto, Ontario.<br />

ALL Consulting. 2003. Handbook on Coalbed Methane Produced Water: Management and Beneficial<br />

Use Alternatives. ALL Consulting for Ground Water Protection Research Foundation, US<br />

Department of Energy, and Bureau of Land Management. Available at http://www.allllc.com/publicdownloads/CBM_BU_Screen.pdf.<br />

Accessed November 30, 2012.<br />

Ambrose, R. B., Hill, S. I. and Mulkey, L. A. 1983. User’s Manual for the Chemical Transport and Fate<br />

Model (TOXIWASP), Version 1. EPA 600/3-83-005. US Environmental Protection Agency.<br />

188 p. Available at http://www.epa.gov/nscep/index.html. Accessed November 30, 2012.<br />

Ambrose, R. B. and Wool, T. A. 2009. WASP7 Stream Transport Model Theory and User’s Guide.<br />

EPA/600/R-09/100. US Environmental Protection Agency. 43 p. Available at<br />

http://www.epa.gov/nscep/index.html. Accessed November 30, 2012.<br />

American Petroleum Institute. 2010. Water Management Associated with Hydraulic Fracturing.<br />

Available at http://www.api.org/Standards/new/api-<strong>hf</strong>2.cfm. Accessed January 20, 2011.<br />

Ashley, G. H. and Robinson, J. F. 1922. The Oil and Gas Fields of Pennsylvania. Mineral Resource<br />

Report M1. Pennsylvania Bureau of Topographic and Geological Survey, Harrisburg,<br />

Pennsylvania.<br />

Agency for Toxic Substances and Disease Registry. 2007. Toxicological Profile for Benzene.<br />

Available at http://www.atsdr.cdc.gov/toxprofiles/tp.aspid=40&tid=14. Accessed<br />

December 4, 2012.<br />

Bradford County Government. Natural Gas Information. Chart of Permitted Gas Wells. Available at<br />

http://bradfordcountypa.org/Natural-Gas.aspspecifTab=2. Accessed August 16, 2012.<br />

Brathwaite, L. D. 2009. Shale-Deposited Natural Gas: A Review of Potential. Presented at California<br />

Energy Commission, Sacramento, California. Available at http://www.energy.ca.gov/<br />

2009publications/CEC-200-2009-005/CEC-200-2009-005-SD.PDF. Accessed November 30,<br />

2012.<br />

172


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Brown, D., Bridgeman, J. and West, J. R. 2011. Predicting chlorine decay and THM formation in<br />

water supply systems. Reviews in Environmental Science and Biotechnology 10 (1): 79-99.<br />

Bruner, K. R. and Smosna, R. 2011. A Comparative Study of the Mississippi Barnett Shale, Fort<br />

Worth Basin, and Devonian Marcellus Shale, Appalachian Basin. DOE/NETL-2011/1478.<br />

URS Corporation for US Department of Energy. Available at http://www.netl.doe.gov/<br />

technologies/oil-gas/publications/brochures/DOE-NETL-2011-1478%20Marcellus­<br />

Barnett.pdf. Accessed December 5, 2012.<br />

Cadmus Group Inc. 2012a. Quality Assurance Project Plan: Modeling the Impact of Hydraulic<br />

Fracturing on Water Resources Based on Water Acquisition Scenarios. Cadmus Group Inc.<br />

for US Environmental Protection Agency. Available at http://www.epa.gov/<strong>hf</strong>study/<br />

qapps.html. Accessed November 30, 2012.<br />

Cadmus Group Inc. 2012b. Supplemental Programmatic Quality Assurance Project Plan: National<br />

Hydraulic Fracturing Study Evaluation of Existing Production Well File Contents. Cadmus<br />

Group Inc. for US Environmental Protection Agency. Available at http://www.epa.gov/<br />

<strong>hf</strong>study/qapps.html. Accessed December 4, 2012.<br />

Cao, J., Li, H., Chow, J. C., Watson, J. G., Lee, S., Rong, B., Dong, J. G. and Ho, K. F. 2011. Chemical<br />

composition of indoor and outdoor atmospheric particles at Emperor Qin's Terra-cotta<br />

Museum. Aerosol and Air Quality Research 11 (1): 70-79.<br />

Cappa, F. and Rutqvist, J. 2011a. Impact of CO 2 geological sequestration on the nucleation of<br />

earthquakes. Geophysical Research Letters 38 (17): L17313.<br />

Cappa, F. and Rutqvist, J. 2011b. Modeling of coupled deformation and permeability evolution<br />

during fault reactivation induced by deep underground injection of CO 2. International<br />

Journal of Greenhouse Gas Control 5 (2): 336-346.<br />

Cappa, F. and Rutqvist, J. 2012. Seismic rupture and ground accelerations induced by CO 2 injection<br />

in the shallow crust. Geophysical Journal International 190 (3): 1784-1789.<br />

Cappa, F., Rutqvist, J. and Yamamoto, K. 2009. Modeling crustal deformation and rupture processes<br />

related to upwelling of deep CO 2-rich fluids during the 1965–1967 Matsushiro earthquake<br />

swarm in Japan. Journal of Geophysical Research 114 (10): B10304.<br />

Carlson, C. G. 1985. Geology of McKenzie County, North Dakota. North Dakota Geological Survey<br />

Bulletin 80 Part 1. North Dakota Geological Survey in cooperation with the US Geological<br />

Survey, North Dakota State Water Commission, and the McKenzie County Water<br />

Management District. 54 p. Available at http://www.swc.state.nd.us/4dlink9/4dcgi/<br />

GetSubContentPDF/PB-285/McKenzie_Part_1.pdf. Accessed December 5, 2012.<br />

Carter, K. M. and Harper, J. A. 2002. Oil and Gas Prospects in Northeastern Pennsylvania. In JD<br />

Inners and DD Braun, editors, From Tunkhannock to Starrucca: Bluestone, Glacial Lakes,<br />

and Great Bridges in the "Endless Mountains" of Northeastern Pennsylvania. Guidebook for<br />

173


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

the 67th Annual Field Conference of Pennsylvania Geologists. Field Conference of<br />

Pennsylvania Geologists, Tunkhannock, Pennsylvania. pp. 15-31.<br />

CDM and GBSM. 2004. Statewide Water Supply Initiative Report. CDM and GBSM for Colorado<br />

Water Conservation Board. Available at http://cospl.coalliance.org/fedora/repository/<br />

co:3490. Accessed December 3, 2012.<br />

Colorado Department of Public Health and Environment. 2012. Colorado’s Section 303(D) List of<br />

Impaired Water and Monitoring and Evaluation List. Available at<br />

http://www.colorado.gov/cs/Satelliteblobcol=urldata&blobheadername1=Content­<br />

Disposition&blobheadername2=Content­<br />

Type&blobheadervalue1=inline%3B+filename%3D%22Section+303(d)+List+and+Colorad<br />

o+Monitoring+and+Evaluation+List+(Regulation+%2393).pdf%22&blobheadervalue2=app<br />

lication%2Fpdf&blobkey=id&blobtable=MungoBlobs&blobwhere=1251806966544&ssbina<br />

ry=true. Accessed November 27, 2012.<br />

Chapman, E. C., Capo, R. C., Stewart, B. W., Kirby, C.S., , Hammack, R. W., Schroeder, K. T. and<br />

Edenborn, H. M. 2012. Geochemical and strontium isotope characterization of produced<br />

water from Marcellus shale natural gas extraction. Environmental Science & Technology 46<br />

(6): 3545-3553.<br />

Molecular Operating Environment (MOE) Linux version 2011.10. 2011. Chemical Computing Group,<br />

Montreal, Quebec.<br />

Colorado Oil and Gas Conservation Commission. 2011. Oil and Gas Industry Spills and Releases.<br />

Available at http://dnr.state.co.us/SiteCollectionDocuments/SpillsAndReleases.pdf.<br />

Accessed December 4, 2012.<br />

Colorado Oil and Gas Conservation Commission. 2012a. Fact Sheet: Water Sources and Demand for<br />

the Hydraulic Fracturing of Oil and Gas Wells in Colorado from 2010 through 2015.<br />

Available at http://cogcc.state.co.us/Library/Oil_and_Gas_Water_Sources_Fact_Sheet.pdf.<br />

Accessed December 12, 2012.<br />

Colorado Oil and Gas Conservation Commission. 2012b. Production Inquiry. Colorado Oil and Gas<br />

Information System. Available at http://cogcc.state.co.us. Accessed October 17, 2012.<br />

Colborn, T., Kwiatkowski, C., Schultz, K. and Bachran, M. 2011. Natural Gas Operations from a Public<br />

Health Perspective. Human and Ecological Risk Assessment 17 (5): 1039-1056.<br />

Coleman, J., Milici, R., Cook, T., Charpentier, R., Kirschbaum, M., Klett, T., Pollastro, R. and Schenk, C.<br />

2011. Assessment of Undiscovered Oil and Gas Resources of the Devonian Marcellus Shale<br />

of the Appalachian Basin Province. National Assessment of Oil and Gas Fact Sheet 2011­<br />

3092. US Geological Survey. 2 p. Available at http://pubs.usgs.gov/fs/2011/3092/.<br />

Accessed November 30, 2012.<br />

174


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Cooper, J. R., Crelling, J. C., Rimmer, S. M. and Whittington, A. G. 2007. Coal Metamorphism by<br />

Igneous Intrusion in the Raton Basin, Colorado, and New Mexico: Implications for<br />

Generation of Volatiles. International Journal of Coal Geology 71 (1): 15-27.<br />

Davies, R. J. 2011. Methane contamination of drinking water caused by hydraulic fracturing remains<br />

unproven. Proceedings of the National Academy of Sciences 108 (43): E871.<br />

Deng, Z.-Q., Bengtsson, L., Singh, V. P. and Adrian, D. D. 2002. Longitudinal dispersion coefficient in<br />

single-channel streams. Journal of Hydraulic Engineering 128 (10): 901-916.<br />

DiToro, D. M., Fitzpatrick, J. J. and Thomann, R. V. 1981. Documentation for Water Quality Analysis<br />

Simulation Program (WASP) and Model Verification Program (MVP). EPA 600/3-81-044. US<br />

Environmental Protection Agency, Duluth, Minnesota.<br />

Dolly, E. D. and Meissner, F. F. 1977. Geology and Gas Exploration Potential, Upper Cretaceous and<br />

Lower Tertiary Strata, Northern Raton Basin, Colorado. In H.K. Veal, editors, Exploration<br />

Frontiers of the Central and Southern Rockies: Rocky Mountain Association of Geologists<br />

Guidebook. Rocky Mountain Association of Geologists, Denver, Colorado. pp. 247-270.<br />

Donigian Jr., A. S. Bibliography for HSPF and Related References. Available at<br />

http://www.aquaterra.com/resources/hspfsupport/hspfbib.php. Accessed<br />

Donigian Jr., A. S., Bicknell, B. and Bandurraga, M. 2011. Watershed Modeling for the Santa Clara<br />

River in Southern California. ASCE EWRI Conference Proceedings on CD-ROM: Bearing<br />

Knowledge for Sustainability, World Environmental and Water Resources Congress, Palm<br />

Springs, California. Available at http://www.aquaterra.com/resources/pubs/index.php.<br />

Accessed November 21, 2012.<br />

Eastern Research Group Inc. 2011. Final Quality Assurance Project Plan for the Evaluation of<br />

Information on Hydraulic Fracturing. Eastern Research Group Inc. for US Environmental<br />

Protection Agency. Available at http://www.epa.gov/<strong>hf</strong>study/qapps.html. Accessed<br />

December 11, 2012.<br />

Engelder, T. 2012. Peer review letter for Warner, N.R., Jackson, R.B., Darrah, T.H., Osborn, S.G.,<br />

Down, A., Zhao, K., White, A., and Vengosh, A. (2012), Geochemical Evidence for Possible<br />

Natural Migration of Marcellus Formation Brine to Shallow Aquifers in Pennsylvania.<br />

Proceedings of the National Academy of Sciences 109 (30): 11961-11966.<br />

ESRI. 2010a. US Counties Shapefile. ESRI Data & Maps Series. ESRI, Redlands, California.<br />

ESRI. 2010b. US States Shapefile. ESRI Data & Maps Series. ESRI, Redlands, California.<br />

ESRI. 2012. US Major Water Shapefile. ESRI Data & Maps Series. ESRI, Redlands, California.<br />

Farhadian, M., Vachelard, C., Duchez, D. and Larroche, C. 2008. In situ bioremediation of<br />

monoaromatic pollutants in groundwater: A review. Bioresource Technology 99 (13): 5296–<br />

5308.<br />

175


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Flores, R. M. 1993. Geologic and geomorphic controls of coal developments in some tertiary rockymountain<br />

basins, USA. International Journal of Coal Geology 23 (1-4): 43-73.<br />

Flores, R. M. and Bader, L. R. 1999. A Summary of Tertiary Coal Resources of the Raton Basin,<br />

Colorado and New Mexico. In Resource Assessment of Tertiary Coalbeds and Zones in the<br />

Northern Rocky Mountains and Great Plains Region. Professional Paper 1625-A. US<br />

Geological Survey. 35 p. Available at http://pubs.usgs.gov/pp/p1625a/. Accessed December<br />

5, 2012.<br />

Freeman, C. M. 2010. Study of Flow Regimes in Multiply-Fractured Horizontal Wells in Tight Gas<br />

and Shale Gas Reservoir Systems. MS Thesis. Texas A & M University, College Station, Texas.<br />

Freeman, C. M., Moridis, G. J. and Blasingame, T. A. 2011. A numerical study of microscale flow<br />

behavior in tight gas and shale gas reservoir systems. Transport in Porous Media 90 (1):<br />

253-268.<br />

Freeman, C. M., Moridis, G. J., Ilk, D. and Blasingame, T. A. 2009a. A Numerical Study of Microscale<br />

Flow Behavior in Tight Gas and Shale Gas Reservoir Systems. Proceedings of the 2009<br />

TOUGH Symposium, Berkeley, California. Available at http://www.netl.doe.gov/kmd/<br />

RPSEA_Project_Outreach/07122-23%20-%20TOUGH%202009%20Symposium%20-%209­<br />

14-09.pdf. Accessed November 30, 2012.<br />

Freeman, C. M., Moridis, G. J., Ilk, D. and Blasingame, T. A. 2009b. A Numerical Study of Performance<br />

for Tight Gas and Shale Gas Reservoir Systems. Presented at SPE Annual Technical<br />

Conference and Exhibition, New Orleans, Louisiana. Available at http://www.onepetro.org/<br />

mslib/servlet/onepetropreviewid=SPE-124961-MS. Accessed November 30, 2012.<br />

Freeman, C. M., Moridis, G. J., Michael, G. E. and Blasingame, T. A. 2012. Measurement, Modeling, and<br />

Diagnostics of Flowing Gas Composition Changes in Shale Gas Wells. Presented at SPE Latin<br />

American and Caribbean Petroleum Engineering Conference, Mexico City, Mexico. Available<br />

at http://www.onepetro.org/mslib/servlet/onepetropreviewid=SPE-153391-MS.<br />

Accessed November 30, 2012.<br />

Gassman, P. W., Reyes, M. R., Green, C. H. and Arnold, J. G. 2007. The soil and water assessment tool:<br />

historical development, applications, and future research directions. Transactions of the<br />

American Society of Aricultural and Biological Engineers 50 (4): 1211-1250.<br />

Greg Lewicki & Associates. 2001. Raton Basin Coal Mine Feature Inventory. Greg Lewicki &<br />

Associates for Colorado Oil and Gas Conservation Commission. Available at<br />

http://cospl.coalliance.org/fedora/repository/co:3284. Accessed December 3, 2012.<br />

Ground Water Protection Council. 2012a. Disclosures Reported by Month in Frac Focus.<br />

Presentation. Ground Water Protection Council, Oklahoma City, Oklahoma.<br />

Ground Water Protection Council. 2012b. FracFocus well records: January 1, 2011, through<br />

February 27, 2012. Available at http://www.fracfocus.org/.<br />

176


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Ground Water Protection Council and ALL Consulting. 2009. Modern Shale Gas Development in the<br />

US: A Primer. Ground Water Protection Council and ALL Consulting for US Department of<br />

Energy. Available at http://www.netl.doe.gov/technologies/oilgas/publications/epreports/shale_gas_primer_2009.pdf.<br />

Accessed December 12, 2012.<br />

Hairer, E., Norsett, S. P. and Wanner, G. 1991. Solving Ordinary Differential Equations I: Nonstiff<br />

Problems. Springer-Verlag, Berlin, Germany.<br />

Haritash, A. K. and Kaushik, C. P. 2009. Biodegradation aspects of polycyclic aromatic hydrocarbons<br />

(PAHs): A review. Journal of Hazardous Materials 169 (1-3): 1-15.<br />

Harrison, S. S. 1983. Evaluating system for ground-water contamination hazards due to gas-well<br />

drilling on the glaciated Appalachian Plateau. Ground Water 21 (6): 689-700.<br />

Harrison, S. S. 1985. Contamination of aquifers by overpressurizing the annulus of oil and gas wells.<br />

Ground Water 23 (7): 317-324.<br />

Hayes, T. 2009. Sampling and Analysis of Water Streams Associated with the Development of<br />

Marcellus Shale Gas. Gas Technology Institute for Marcellus Shale Coalition. Available at<br />

http://eidmarcellus.org/wp-content/uploads/2012/11/MSCommission-Report.pdf.<br />

Accessed November 30, 2012.<br />

Healy, R. W., Bartos, T. T., Rice, C. A., McKinley, M. P. and Smith, B. D. 2011. Groundwater chemistry<br />

near an impoundment for produced water, Powder River Basin, Wyoming USA. Journal of<br />

Hydrology 403 (1-2): 37-48.<br />

Healy, R. W., Rice, C. A., Bartos, T. T. and McKinley, M. P. 2008. Infiltration from an impoundment for<br />

coal-bed natural gas, Powder River Basin, Wyoming: Evolution of water and sediment<br />

chemistry. Water Resources Research 44 (6): W06424.<br />

Holditch, S. A. 1993. Completion methods in coal-seam reservoirs. Journal of Petroleum Technology<br />

45 (3): 270-276.<br />

Homer, C., Dewitz, J., Fry, J., Coan, M., Hossain, N., Larson, C., Herold, N., McKerrow, A., VanDriel, J. N.<br />

and Wickham, J. 2007. Completion of the 2001 national land cover database for the<br />

conterminous United States. Photogrammetric Engineering & Remote Sensing 73 (4): 337­<br />

341.<br />

Howard, P. H. 1989. Handbook of Environmental Fate and Exposure Data for Organic Chemicals.<br />

CRC Press, Syracuse, New York.<br />

Howard, P. H., Boethling, R. S., Jarvis, W. F., Meylan, W. M. and Michalenko, E. M. 1991. Handbook of<br />

Environmental Degradation Rates. CRC Press, Boca Raton, Florida.<br />

Ivahnenko, T. and Flynn, J. L. 2010. Estimated Withdrawals and Use of Water in Colorado, 2005.<br />

Scientific Investigations Report 2010-5002. US Geological Survey. 61 p. Available at<br />

http://pubs.usgs.gov/sir/2010/5002/. Accessed November 30, 2012.<br />

177


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Jackson, R. B., Osborn, S. G., Vengosh, A. and Warner, N. 2011. Reply to Davies: Hydraulic fracturing<br />

remains a possible mechanism for observed methane contamination of drinking water.<br />

Proceedings of the National Academy of Sciences 108 (43): E872.<br />

Jacob, R. 2011. Incident Action Plan, Franchuk 44-20 SWH Incident. Denbury Onshore, LLC, Plano,<br />

Texas.<br />

Jeu, S. J., Logan, T. L. and McBane, R. A. 1988. Exploitation of Deeply Buried Coalbed Methane Using<br />

Different Hydraulic Fracturing Techniques in the Piceance Basin, Colorado, and San Juan<br />

Basin, New Mexico. Presented at SPE Annual Technical Conference and Exhibition, Houston,<br />

Texas. Available at http://www.onepetro.org/mslib/app/<br />

Preview.dopaperNumber=00018253&societyCode=SPE. Accessed November 30, 2012.<br />

Jobson, H. E. 1996. Prediction of Traveltime and Longitudinal Dispersion in Rivers and Streams.<br />

Water-Resources Investigations Report 96-4013. US Geological Survey. 72 p. Available at<br />

http://water.usgs.gov/osw/pubs/wrir964013header.html. Accessed November 30, 2012.<br />

Johnson, N., Gagnolet, T., Ralls, R., Zimmerman, E., Eichelberger, B., Tracey, C., Kreitler, G., Orndorff,<br />

S., Tomlinson, J., Bearer, S. and Sargent, S. 2010. Marcellus Shale Natural Gas and Wind.<br />

Pennsylvania Energy Impacts Assessment Report 1. The Nature Conservancy. 47 p.<br />

Available at http://www.nature.org/media/pa/tnc_energy_analysis.pdf. Accessed<br />

November 30, 2012.<br />

Johnson, R. C. and Finn, T. M. 2001. Potential for a Basin-Centered Gas Accumulation in the Raton<br />

Basin, Colorado and New Mexico. Bulletin 2184-B. US Geological Survey. 18 p. Available at<br />

http://pubs.usgs.gov/bul/b2184-b/b2184-b.pdf. Accessed December 3, 2012.<br />

Johnson, T. E., Butcher, J. B., Parker, A. and Weaver, C. P. 2011. Investigating the sensitivity of U.S.<br />

streamflow and water quality to climate change: the U.S. EPA global change research<br />

program’s “20 watersheds” project. Journal of Water Resources Planning and Management<br />

138 (5): 453-464.<br />

Jurich, A. and Adams, M. A. 1984. Geologic Overview, Coal, and Coalbed Methane Resources of Raton<br />

Mesa Region – Colorado and New Mexico. In C.T. Rightmire, G.E. Eddy and J.N. Kirr, editors,<br />

Studies in Geology Series 17: Coalbed Methane Resources of the United States. American<br />

Association of Petroleum Geologists, Tulsa, Oklahoma. pp. 163–184.<br />

Keighin, C. W. 1995. Raton Basin-Sierra Grande Uplift Province (041). National Assessment of<br />

United States Oil and Gas Resources-Results, Methodology, and Supporting Data. Digital<br />

Data Series DDS-30. US Geological Survey. 1-7 p. Available at<br />

http://certmapper.cr.usgs.gov/data/noga95/prov41/text/prov41.pdf. Accessed December<br />

3, 2012.<br />

Kenny, J. F., Barter, N. L., Hutson, S. S., Linsey, K. S., Lovelace, J. K. and Maupin, M. A. 2009. Estimated<br />

Use of Water in the United States in 2005. Circular 1344. US Geological Survey. p. Available<br />

at http://pubs.usgs.gov/circ/1344/. Accessed November 30, 2012.<br />

178


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Kim, J. and Moridis, G. J. 2012a. Analysis of fracture propagation during hydraulic fracturing<br />

operations in tight/shale gas systems. In preparation for journal submission. Lawrence<br />

Berkeley National Laboratory, Berkeley, California.<br />

Kim, J. and Moridis, G. J. 2012b. Development of the T+M Coupled Flow-Geomechanical Simulator to<br />

Describe Fracture Propagation and Coupled Flow-Thermal-Geomechanical Processes in<br />

Tight/Shale Gas Systems. Proceedings of the 2012 TOUGH Symposium Berkeley, California.<br />

Available at http://esd.lbl.gov/files/research/projects/tough/events/symposia/<br />

toughsymposium12/Kim_Jihoon-T+M.pdf. Accessed November 30, 2012.<br />

Kim, J. and Moridis, G. J. 2012c. Gas Flow Tightly Coupled to Elastoplastic Geomechanics for Tight<br />

and Shale Gas Reservoirs: Material Failure and Enhanced Permeability. Presented at SPE<br />

Americas Unconventional Resources Conference, Pittsburgh, Pennsylvania. Available at<br />

http://www.onepetro.org/mslib/app/Preview.dopaperNumber=SPE-155640­<br />

MS&societyCode=SPE. Accessed November 30, 2012.<br />

Kim, J. and Moridis, G. J. 2012d. Numerical Geomechanical Analyses on Hydraulic Fracturing in Tight<br />

Gas and Shale Gas Reservoirs. Presented at 2012 SPE Annual Technical Conference and<br />

Exhibition, San Antonio, Texas.<br />

Kim, J. and Moridis, G. J. 2012e. Numerical Studies for Naturally Fractured Shale Gas Reservoirs:<br />

Coupled Flow and Geomechanics in Multiple Porosity/Permeability Materials. Presented at<br />

46th U.S. Rock Mechanics/Geomechanics Symposium, Chicago, Illinois.<br />

Kish, L. 1965. Survey Sampling. John Wiley & Sons, New York, New York.<br />

Klausing, R. L. 1979. Ground-Water Resources of Dunn County, North Dakota. US Geological Survey<br />

for North Dakota State Water Commission and North Dakota Geologic Survey. Available at<br />

http://www.swc.state.nd.us/4dlink9/4dcgi/GetSubContentPDF/PB-225/Dunn_Part_3.pdf.<br />

Accessed December 3, 2012.<br />

Kuuskraa, V. A. and Ammer, J. 2004. Tight gas sands development–how to dramatically improve<br />

recovery efficiency. GasTIPS 10 (1): 15-20.<br />

Lawrence Berkeley National Laboratory. 2011. Quality Assurance Project Plan: Analysis of<br />

Environmental Hazards Related to Hydrofracturing. Lawrence Berkeley National<br />

Laboratory for US Environmental Protection Agency. Available at http://www.epa.gov/<br />

<strong>hf</strong>study/qapps.html. Accessed November 27, 2012.<br />

Leadscope version 3.0.6-3. 2012. Leadscope Inc., Columbus, Ohio.<br />

Lee, J. J. 2011. Hydraulic Fracturing and Safe Drinking Water. Proceedings of the US Environmental<br />

Protection Agency Technical Workshops for the Hydraulic Fracturing Study: Water<br />

Resources Management, Arlington, Virginia. Available at http://epa.gov/<strong>hf</strong>study/<br />

waterworkshop.html. Accessed November 30, 2012.<br />

179


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Leonard, B. P. 2002. Stability of explicit advection schemes. The balance point location rule.<br />

International Journal for Numerical Methods in Fluids 38 (5): 471-514.<br />

Lide, D. R. 2008. CRC Handbook of Chemistry and Physics. 89th Edition. CRC Press, Boca Raton,<br />

Florida.<br />

Macartney, H. 2011. Hydraulic Fracturing in Coalbed Methane Development, Raton Basin, Southern<br />

Colorado. Presented at US Environmental Protection Agency Technical Workshop for the<br />

Hydraulic Fracturing Study: Well Construction and Operations, Arlington, Virginia. Available<br />

at http://www.epa.gov/<strong>hf</strong>study/wellconstructworkshop.html. Accessed December 5, 2012.<br />

Marcellus Center for Outreach and Research. 2012a. Depth of Marcellus Shale Base. Pennsylvania<br />

State University. Available at http://www.marcellus.psu.edu/images/Marcellus_Depth.gif.<br />

Accessed December 6, 2012.<br />

Marcellus Center for Outreach and Research. 2012b. Extent and Thickness of Marcellus Shale.<br />

Pennsylvania State University. Available at http://www.marcellus.psu.edu/images/<br />

Marcellus_thickness.gif. Accessed December 6, 2012.<br />

MarcellusGas.Org. 2012a. Estimated Anticipated Water Usage (for Hydraulic Fracturing) for<br />

Permitted Horizontal Wells, by Township, for Washington County, Pennsylvania. Available<br />

at http://www.marcellusgas.org/graphs/PA-Washington#waterstat. Accessed September<br />

27, 2012.<br />

MarcellusGas.Org. 2012b. Marcellus Well Permits, by Year, for Washington County, Pennsylvania.<br />

Available at http://www.marcellusgas.org/graphs/PA-Washington#pyear. Accessed<br />

September 25, 2012.<br />

MarcellusGas.Org. 2012c. Marcellus Well Starts, by Year, for Washington County, Pennsylvania.<br />

Available at http://www.marcellusgas.org/graphs/PA-Washington#wstarts. Accessed<br />

September 25, 2012.<br />

Material Safety Data Sheets. (a) Encana/Halliburton Energy Services, Inc.: Duncan, Oklahoma.<br />

Provided by Halliburton Energy Services during an onsite visit by the EPA on May 10, 2010;<br />

(b) Encana Oil and Gas (USA), Inc.: Denver, Colorado. Provided to US EPA Region 8.<br />

Mazzoldi, A., Rinaldi, A. P., Borgia, A. and Rutqvist, J. 2012. Induced seismicity within geological<br />

carbon sequestration projects: Maximum earthquake magnitude and leakage potential from<br />

undetected faults. International Journal of Greenhouse Gas Control 10 (1): 434-442.<br />

McCarthy. 2009. Travel Times, Streamflow Velocities, and Dispersion Rates in the Yellowstone<br />

River, Montana. Scientific Investigations Report 2009-5261. US Geological Survey. 25 p.<br />

Available at http://pubs.usgs.gov/sir/2009/5261/. Accessed November 30, 2012.<br />

McGonigal, K. H. 2005. Nutrients and Suspended Sediment Transported in the Susquehanna River<br />

Basin, 2004, and Trends, January 1985 Through December 2004. Publication 241.<br />

180


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Susquehanna River Basin Commission. 61 p. Available at http://www.srbc.net/<br />

pubinfo/techdocs/Publication_241/NutrientReport.pdf. Accessed November 30, 2012.<br />

Milici, R. C. and Swezey, C. S. 2006. Assessment of Appalachian Basin Oil and Gas Resources:<br />

Devonian Shale–Middle and Upper Paleozoic Total Petroleum System. Open-File Report<br />

Series 2006-1237. US Geological Survey. 70 p. Available at http://pubs.usgs.gov/of/2006/<br />

1237/of2006-1237.pdf. Accessed December 4, 2012.<br />

Montgomery, J. H. 2000. Groundwater Chemicals Desk Reference. 3rd Edition. CRC Press, Boca<br />

Raton, Florida.<br />

Moridis, F. and Webb, F. 2012. The RealGas and RealGasH2O Options of the TOUGH+ Code for the<br />

Simulation of Coupled Fluid and Heat Flow in Tight/Shale Gas Systems Proceedings of the<br />

2012 TOUGH Symposium, Berkeley, California. Available at http://esd.lbl.gov/files/<br />

research/projects/tough/events/symposia/toughsymposium12/Moridis_George­<br />

RealgasH2O.pdf. Accessed November 30, 2012.<br />

Moridis, G. J., Blasingame, T. and Freeman, C. M. 2010. Analysis of Mechanisms of Flow in Fractured<br />

Tight-Gas and Shale-Gas Reservoirs. Presented at 2010 SPE Latin American & Caribbean<br />

Petroleum Engineering Conference, Lima, Peru. Available at http://www.onepetro.org/<br />

mslib/servlet/onepetropreviewid=SPE-139250-MS. Accessed November 30, 2012.<br />

Moridis, G. J. and Freeman, C. M. 2012. The RGasH2OCont module of the TOUGH+ code for<br />

simulation of coupled fluid and heat flow, and contaminant transport, in tight/shale gas<br />

systems. In preparation for journal submission. Lawrence Berkeley National Laboratory,<br />

Berkely, California.<br />

Moridis, G. J., Kowalsky, M. B. and Pruess, K. 2008. TOUGH+HYDRATE v1.0 User Manual: A Code for<br />

the Simulation of System Behavior in Hydrate-Bearing Geologic Media. Report 149E.<br />

Lawrence Berkley National Laboratory. 279 p. Available at http://esd.lbl.gov/files/<br />

research/projects/tough/documentation/TplusH_Manual_v1.pdf. Accessed November 30,<br />

2012.<br />

Murphy, E. C. 2001. Geology of Dunn County. Bulletin 68 Part 1. North Dakota Geological Survey. 43<br />

p. Available at http://www.swc.state.nd.us/4dlink9/4dcgi/GetSubContentPDF/PB­<br />

221/Dunn_Part_1.pdf. Accessed December 4, 2012.<br />

Myers, T. 2012a. Author’s Reply. Ground Water 50 (6): 828-830.<br />

Myers, T. 2012b. Potential contaminant pathways from hydraulically fractured shale to aquifers.<br />

Ground Water 50 (6): 872-882.<br />

National Academy of Sciences. 2012. Induced Seismicity Potential in Energy Technologies. Available<br />

at http://dels.nas.edu/besr. Accessed November 30, 2012.<br />

North Dakota Industrial Commission. 2003. January 2003 Oil and Gas Production Report. Available<br />

at https://www.dmr.nd.gov/oilgas/mpr/2003_01.pdf. Accessed November 27, 2012.<br />

181


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

North Dakota Industrial Commission. 2012a. North Dakota Monthly Bakken Oil Production<br />

Statistics Available at https://www.dmr.nd.gov/oilgas/stats/historicalbakkenoilstats.pdf.<br />

Accessed November 27, 2012.<br />

North Dakota Industrial Commission. 2012b. Oil and Gas Production Report. Available at<br />

https://www.dmr.nd.gov/oilgas/mpr/2012_07.pdf. Accessed December 5, 2012.<br />

Newport, T. G. 1973. Summary Ground-Water Resources of Washington County, Pennsylvania.<br />

Water Resource Report 38. Pennsylvania Geological Survey. 32 p. Available at<br />

http://www.dcnr.state.pa.us/topogeo/pub/water/pdfs/w038_text.pdf. Accessed December<br />

4, 2012.<br />

Nicot, J., Hebel, A., Ritter, S., Walden, S., Baier, R., Galusky, P., Beach, J., Kyle, R., Symank, L. and<br />

Breton, C. 2011. Current and Projected Water Use in the Texas Mining and Oil and Gas<br />

Industry. The University of Texas at Austin Bureau of Economic Geology for Texas Water<br />

Development Board. Available at http://www.twdb.texas.gov/publications/reports/<br />

contracted_reports/doc/0904830939_MiningWaterUse.pdf. Accessed November 10, 2012.<br />

Noorishad, J. and Tsang, C.-F. 1997. Coupled Thermohydroelasticity Phenomena in Variably<br />

Saturated Fractured Porous Rocks - Formulation and Numerical Solution. In O. Stephanson,<br />

L. Jing and C.-F. Tsang, editors, Coupled Thermo-Hydro-Mechanical Processes in Fractured<br />

Media: Mathematical and Experimental Studies. Elsevier Science B.V., Amsterdam, The<br />

Netherlands. pp. 575.<br />

Nordeng, SH. 2010. The Bakken Source System: Emphasis on the Three Forks Formation. North<br />

Dakota Geological Survey. Available at http://www.ndoil.org/image/cache/<br />

Stephan_Nordeng_-_NDGS.pdf. Accessed December 5, 2012.<br />

Nordin, C. F. and Sabol, G. V. 1974. Empirical Data on Longitudinal Dispersion in Rivers. Water-<br />

Resources Investigations 20-74. US Geological Survey, p. 332.<br />

Nordstrom, P. L. 1982. Occurrence, Availiability and Chemical Quality of Ground Water in the<br />

Cretaceous Aquifers of North-Central Texas. Volume 1. Report 269. Texas Department of<br />

Water Resources. 66 p. Available at http://www.twdb.state.tx.us/publications/reports/<br />

numbered_reports/doc/R269/R269v1/R269v1.pdf. Accessed December 3, 2012.<br />

New York State Department of Environmental Conservation. 2011. Supplemental Generic<br />

Environmental Impact Statement on the Oil, Gas and Solution Mining Regulatory Program<br />

(Revised Draft). Well Permit Issuance for Horizontal Drilling and High-Volume Hydraulic<br />

Fracturing to Develop the Marcellus Shale and Other Low-Permeability Gas Reservoirs.<br />

Available at http://www.dec.ny.gov/energy/75370.html. Accessed September 1, 2011.<br />

Olorode, O. M. 2011. Numerical modeling of fractured shale-gas and tight-gas reservoirs using<br />

unstructured grids. MS Degree. Texas A&M University, College Station, Texas.<br />

182


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Osborn, S. G., Vengosh, A., Warner, N. R. and Jackson, R. B. 2011. Methane contamination of drinking<br />

water accompanying gas-well drilling and hydraulic fracturing. Proceedings of the National<br />

Academy of Sciences 108 (20): 8172-8176.<br />

Pennsylvania Department of Environmental Protection. 2010. Chemicals Used by Hydraulic<br />

Fracturing Companies in Pennsylvania for Surface and Hydraulic Fracturing Activities.<br />

Available at http://files.dep.state.pa.us/OilGas/BOGM/BOGMPortalFiles/MarcellusShale/<br />

Frac%20list%206-30-2010.pdf. Accessed November 27, 2012.<br />

Pennsylvania Department of the Environment. 2011. Letter from Pennsylvania Department of the<br />

Environment to US EPA Region 3 Administrator Shawn Garvin. Available at<br />

http://www.epa.gov/region3/marcellus_shale#inforeqsbypadep. Accessed April 6, 2011.<br />

Pennsylvania Department of Environmental Protection. 2012. State Water Plan Digital Water Atlas.<br />

Available at http://www.pawaterplan.dep.state.pa.us/statewaterplan/DWA/<br />

DWAMain.aspxtheme=6&cacheId=227881113. Accessed August 13, 2012.<br />

Palmer, I. D., Fryar, R. T., Tumino, K. A. and Puri, R. 1991. Water fracs outperform gel fracs in<br />

coalbed pilot. Oil & Gas Journal 89 (32): 71-76.<br />

Palmer, I. D., Lambert, S. W. and Spitler, J. L. 1993. Coalbed Methane Well Completions and<br />

Stimulations: Chapter 14. In Ben E. Law and Dudley D. Rice, editors, SG 38: Hydrocarbons<br />

from Coal. American Association of Petroleum Geologists, Tulsa, Oklahoma. pp. 303-341.<br />

Pancras, J. P., Vedantham, R., Landis, M. S., Norris, G. A. and Ondov, J. M. 2011. Application of EPA<br />

unmix and non-parametric wind regression on high time resolution trace elements and<br />

speciated mercury in Tampa, Florida, aerosol. Environmental Science & Technology 45 (8):<br />

3511-3518.<br />

Plewa, M. J., Muellner, M. G., Richardson, S. D., Fasano, F., Buettner, K. M., Woo, Y.-T., McKague, B. and<br />

Wagner, E. D. 2008. Occurrence, synthesis and mammalian cell cytotoxicity and genotoxicity<br />

of haloacetamides: an emerging class of nitrogenous drinking water disinfection<br />

byproducts. Environmental Science & Technology 42 (3): 955-961.<br />

Plewa, M. J. and Wagner, E. D. 2009. Quantitative Comparative Mammalian Cell Cytotoxicity and<br />

Genotoxicity of Selected Classes of Drinking Water Disinfection By-Products. Presented at<br />

Water Research Foundation, Denver, Colorado. Available at http://prod2010.waterrf.org/<br />

ExecutiveSummaryLibrary/91249_3089_profile.pdf. Accessed November 30, 2012.<br />

Plewa, M. J., Wagner, E. D., Jazwierska, P., Richardson, S. D., Chen, P. H. and McKague, A. B. 2004.<br />

Chemical and biological characterization of newly discovered iodoacid drinking water<br />

disinfection byproducts. Environmental Science & Technology 38 (18): 4713-4722.<br />

Puder, M. and Veil, J. 2006. Offsite Commercial Disposal of Oil and Gas Exploration and Production<br />

Waste: Availability, Options, and Costs. ANL/EVS/R-06/5. Argonne National Laboratory for<br />

US Department of Energy. Available at http://www.evs.anl.gov/pub/doc/ANL-EVS-R-06­<br />

5.pdf. Accessed October 18, 2012.<br />

183


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Reed, L. A. and Stuckey, M. A. 2002. Prediction of Velocities for a Range of Streamflow Conditions in<br />

Pennsylvania. Water-Resources Investigations Report 01-4214. US Geological Survey. 13 p.<br />

Available at http://pubs.er.usgs.gov/publication/wri014214. Accessed November 21, 2012.<br />

Reutter, D. C. and Dunn, D. D. 2000. Water-Quality Assessment of the Trinity River Basin, Texas:<br />

Ground-Water Quality of the Trinity, Carrizo-Wilcox, and Gulf Coast Aquifers, February-<br />

August 1994. Water-Resources Investigations Report 99-4233. US Geological Survey. 64 p.<br />

Available at http://pubs.usgs.gov/wri/wri99-4233/. Accessed December 5, 2012.<br />

Richardson, S. 2003. Disinfection by-products and other emerging contaminants in drinking water.<br />

Trends in Analytical Chemistry 22 (10): 666-684.<br />

Richardson, S. D., Plewa, M. J., Wagner, E. D., Schoeny, R. and DeMarini, D. M. 2007. Occurrence,<br />

genotoxicity, and carcinogenicity of regulated and emerging disinfection by-products in<br />

drinking water: a review and roadmap for research. Mutatation Research 636 (1-3): 178­<br />

242.<br />

Richenderfer, J. 2011. Natural Gas Industry Effects on Water Consumption and Management.<br />

Susquehanna River Basin Commission. 40 p. Available at http://www.mde.state.md.us/<br />

programs/Land/mining/marcellus/Pages/surfacewater.aspx. Accessed November 30,<br />

2012.<br />

Rogers, S. W., Ong, S. K., Kjartanson, B. H., Golchin, J. and Stenback, G. A. 2002. Natural attenuation of<br />

polycyclic aromatic hydrocarbon-contaminated sites: review. Practice Periodical of<br />

Hazardous, Toxic, and Radioactive Waste Management 6 (3): 141-155.<br />

Rossenfoss, S. 2011. From flowback to fracturing: water recycling grows in the marcellus shale.<br />

Journal of Petroleum Technology 63 (7): 48-51.<br />

Railroad Commission of Texas. 2012. Barnett Shale Information. Available at<br />

http://www.rrc.state.tx.us/barnettshale/index.php. Accessed December 5, 2012.<br />

Rupp, B., Appel, K. E. and Gundert-Remy, U. 2010. Chronic oral LOAEL prediction by using a<br />

commercially available computational QSAR tool. Archives of Toxicology 84 (9): 681-688.<br />

Rutqvist, J. 2012. The geomechanics of CO 2 storage in deep sedimentary formations. Geotechnical<br />

and Geological Engineering 30 (3): 525-551.<br />

Rutqvist, J., Birkholzer, J., Cappa, F. and Tsang, C.-F. 2007. Estimating maximum sustainable<br />

injection pressure during geological sequestration of CO2 using coupled flow and<br />

geomechanical fault-slip analysis. Energy Conversion and Management 48 (6): 1798-1807.<br />

Rutqvist, J., Borgesson, L., Chijimatsu, M., Kobayashi, A., Jing, L., Nguyen, T. S., Noorishad, J. and<br />

Tsang, C.-F. 2001. Thermohydromechanics of partially saturated geological media:<br />

governing equations and formulation of four finite element models. International Journal of<br />

Rock Mechanics and Mining Sciences 38 (1): 105-127.<br />

184


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Rutqvist, J., Rinaldi, A. P., Cappa, F. and Moridis, G. J. 2012. Modeling of fault reactivation and<br />

induced seismicity during hydraulic fracturing of shale-gas reservoirs. In preparation for<br />

journal submission. Lawrence Berkeley National Laboratory, Berkeley, California.<br />

S.S. Papadopulos & Associates Inc. 2008. Coalbed Methane Stream Depletion Assessment Study –<br />

Raton Basin, Colorado. S.S. Papadopulos & Associates Inc. prepared in conjunction with the<br />

Colorado Geological Survey for Colorado Department of Natural Resources and the<br />

Colorado Oil and Gas Conservation Commission. Available at http://geosurvey.state.co.us/<br />

water/CBM%20Water%20Depletion/Documents/RatonCBMdepletion_FINAL.pdf. Accessed<br />

December 3, 2012.<br />

S.S. Papadopulos & Associates Inc. 2007a. Coalbed Methane Stream Depletion Assessment Study –<br />

Piceance Basin, Colorado. S.S. Papadopulos & Associates Inc. for Prepared in conjunction<br />

with the Colorado Geological Survey for the State of Colorado Department of Natural<br />

Resources and the Colorado Oil and Gas Conservation Commission. Available at<br />

http://water.state.co.us/groundwater/GWAdmin/NontribGW/Archive/Pages/CBMStream<br />

DepletionStudies.aspx. Accessed November 30, 2012.<br />

S.S. Papadopulos & Associates Inc. 2007b. Coalbed Methane Stream Depletion Assessment Study –<br />

Raton Basin, Colorado, Draft Final Report. S.S. Papadopulos & Associates Inc. for Prepared<br />

in conjunction with the Colorado Geological Survey for the State of Colorado Department of<br />

Natural Resources and the Colorado Oil and Gas Conservation Commission. Available at<br />

http://water.state.co.us/groundwater/GWAdmin/NontribGW/Archive/Pages/CBMStream<br />

DepletionStudies.aspx. Accessed November 30, 2012.<br />

SAIC Energy Environment & Infrastructure LLC and Groundwater & Environmental Services Inc.<br />

2011. ATGAS Investigation Initial Site Characterization and Response, April 19, 2011 to May<br />

2, 2011, ATGAS2H Well Pad, Permit No. 37-015-21237, Leroy Township, Bradford County,<br />

PA. SAIC and GES for Chesapeake Appalachia, LLC. Available at http://www.chk.com/news/<br />

articles/documents/atgas_initial_site_characterization_report_final_08292011.pdf.<br />

Accessed December 4, 2012.<br />

Satterfield, J., Kathol, D., Mantell, M., Hiebert, F., Lee, R. and Patterson, K. 2008. Managing Water<br />

Resource Challenges in Select Natural Gas Shale Plays. Presented at Ground Water<br />

Protection Council Annual Forum, Oklahoma City, Oklahoma.<br />

Schiesser, W. E. 1991. The Numerical Method-of-Lines: Integration of Partial Differential Equations.<br />

Academic Press, San Diego, California.<br />

Qikprop version 3.4. 2012. Schrodinger, LLC, New York, New York.<br />

Schwarzenbach, R. P., Gschwend, P. M. and Imboden, D. M. 2002. Environmental Organic Chemistry.<br />

2nd Edition. John Wiley & Sons, Inc, Hoboken, New Jersey.<br />

Scott, G. and Armstrong, J. M. 1932. The Geology of Wise County, Texas. The University of Texas<br />

Bulletin No. 3224. The University of Texas. 80 p. Available at http://www.lib.utexas.edu/<br />

185


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

books/landscapes/publications/txu-oclc-983185/txu-oclc-983185.pdf. Accessed December<br />

5, 2012.<br />

Scott, M. J. and Jones, M. N. 2000. The biodegradation of surfactants in the environment. Biochimica<br />

et Biophysica Acta 1508 (1-2): 235-251.<br />

Seagren, E. A. and Becker, J. G. 2002. Review of natural attenuation of BTEX and MTBE in<br />

groundwater. Practice Periodical of Hazardous, Toxic, and Radioactive Waste Management 6<br />

(3): 156-172.<br />

Seo, J. S., Keum, Y. S. and Qing, X. L. 2009. Bacterial degradation of aromatic compounds.<br />

International Journal Environmental Research and Public Health 6 (1): 278- 309.<br />

Sharma, V. K., Anquandah, G. A. K., Yngard, R. A., Kim, H., Fekete J., Bouzek, K., Ray, A. K. and Golovko,<br />

D. 2009. Nonylphenol, octylphenol, and bisphenol-A in the aquatic environment: a review<br />

on occurrence, fate, and treatment. Journal of Environmental Science and Health, Part A:<br />

Toxic/Hazardous Substances and Environmental Engineering 44 (5): 423-442.<br />

Shultz, C. H. 1999. The Geology of Pennsylvania. Pennsylvania Geological Survey and Pittsburgh<br />

Geological Society, Harrisburg, Pennsylvania.<br />

Soares A., Guieysse, B., Jeffereson, B., Cartmell, E. and J.N., L. 2008. Nonylphenol in the environment:<br />

a critical review on occurrence, fate, toxicity and treatment in wastewaters. Environment<br />

International 34 (7): 1033-1049.<br />

Soonthornnonda, P. and Christensen, E. 2008. Source apportionment of pollutants and flows of<br />

combined sewer wastewater. Water Research 42 (8-9): 1989-1998<br />

Spahr, N. E., Apodaca, L. E., Deacon, J. R., Bails, J. B., Bauch, N. J., Smiath, C. M. and Driver, N. E. 2000.<br />

Water Quality in the Upper Colorado River Basin, Colorado, 1996-98. Circular 1214. US<br />

Geological Survey. 33 p. Available at http://pubs.water.usgs.gov/circ1214/. Accessed<br />

November 30, 2012.<br />

Susquehanna River Basin Commission. Susquehanna River Basin Commission Information Sheet.<br />

Available at http://www.srbc.net/pubinfo/docs/<br />

Susq%20River%20Basin%20General%20%2811_06%29.PDF. Accessed November 27,<br />

2012.<br />

Susquehanna River Basin Commission. 2012. Flowback and Produced Water Volume. Data Provided<br />

to EPA Upon Request. Susquehanna River Basin Commission, Harrisburg, Pennsylvania.<br />

Stangroom, S. J., Collins C.D. and Lester, J. N. 2010. Abiotic behavior of organic micropollutants in<br />

soils and the aquatic environment. A review: II. Transformations. Environmental<br />

Technology 21 (8): 865-882.<br />

Staples, C. A., Williams, J. B., Craig, G. R. and Roberts., K. M. 2001. Fate, effects and potential<br />

environmental risks of ethylene glycol: a review. Chemosphere 43 (3): 377-383.<br />

186


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Talmage, S. S. 1994. Environmental and Human Safety of Major Surfactants—Alcohol Ethoxylates<br />

and Alkylphenol Ethoxylates. CRC Press, Boca Raton, Florida.<br />

Taylor, L. E. 1984. Groundwater Resources of the Upper Susquehanna River Basin, Pennsylvania.<br />

Pennsylvania Bureau of Topographic and Geologic Survey, Harrisburg, Pennsylvania.<br />

Topper, R., Scott, K. and Watterson, N. 2011. Geologic Model of the Purgatoire River Watershed<br />

within the Raton Basin, Colorado. Colorado Geological Survey for Colorado Water<br />

Conservation Board. Available at http://geosurvey.state.co.us/water/<br />

CBM%20Water%20Depletion/Documents/Raton%20Geology_CGS%20final%20report.pdf.<br />

Accessed December 3, 2012.<br />

Tremain, C. M. 1980. The Coalbed Methane Potential of the Raton Basin, Colorado. Presented at SPE<br />

Unconventional Gas Recovery Symposium, Pittsburgh, Pennsylvania. Available at<br />

http://www.onepetro.org/mslib/servlet/onepetropreviewid=00008927. Accessed<br />

December 3, 2012.<br />

Tyler, R. 1995. Geologic and Hydrologic Assessment of Natural Gas from Coal: Greater Green River,<br />

Piceance, Powder River, and Raton Bains, Western United States. The University of Texas at<br />

Austin, Bureau of Economic Geology, Austin, Texas.<br />

Tyler, R. and McMurry, R. G. 1995. Genetic Stratigraphy, Coal Occurrence, and Regional Cross<br />

Section of the Williams Fork Formation. Open File Report 95-2. Colorado Geological Survey.<br />

42 p. Available at http://geosurveystore.state.co.us/p-764-genetic-stratigraphy-coaloccurrence-and-regional-cross-section-of-the-coal-bearing-williams-fork-formation.aspx.<br />

Accessed November 30, 2012.<br />

US Energy Information Administration. 2011a. Annual Energy Review 2010. Available at<br />

http://www.eia.gov/aer. Accessed September 26, 2012.<br />

US Energy Information Administration. 2011b. Distribution and Production of Oil and Gas Wells by<br />

State, 1995 to 2009. Available at http://www.eia.gov/pub/oil_gas/petrosystem/<br />

petrosysog.html. Accessed November 27, 2012.<br />

US Energy Information Administration. 2011c. Shale Gas and Oil Plays, Lower 48 States. Available at<br />

http://www.eia.gov/pub/oil_gas/natural_gas/analysis_publications/maps/maps.htm.<br />

Accessed December 12, 2012.<br />

US Energy Information Administration. 2011d. Shapefiles for Basin boundaries. Data for the US<br />

Shale Plays Map. May 9, 2011. US Energy Information Administration, Washington, DC.<br />

Available at http://www.eia.gov/pub/oil_gas/natural_gas/analysis_publications/<br />

maps/maps.htm. Accessed November 30, 2012.<br />

US Energy Information Administration. 2011e. Shapefiles for Play Boundaries. Data for the US Shale<br />

Plays Map. May 9, 2011. US Energy Information Administration, Washington, DC. Available<br />

at http://www.eia.gov/pub/oil_gas/natural_gas/analysis_publications/maps/maps.htm.<br />

Accessed November 30, 2012.<br />

187


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

US Energy Information Administration. 2012. Annual Energy Outlook 2012. Available at<br />

http://www.eia.gov/forecasts/aeo/index.cfm. Accessed November 27, 2012.<br />

US Environmental Protection Agency. 1995. Guidance for Methods Development and Methods<br />

Validation for the Resource Conservation and Recovery Act (RCRA) Program. Available at<br />

http://www.epa.gov/sw-846/pdfs/methdev.pdf. Accessed December 12, 2012.<br />

US Environmental Protection Agency. 2000a. EPA Quality Manual for Environmental Programs.<br />

Available at http://www.epa.gov/irmpoli8/policies/2105P010.pdf. Accessed December 3,<br />

2012.<br />

US Environmental Protection Agency. 2000b. Guidance on Technical Audits and Related<br />

Assessments for Environmental Data Operations. Available at http://www.epa.gov/<br />

quality/qa_docs.html. Accessed December 3, 2012.<br />

US Environmental Protection Agency. 2000c. Policy and Program Requirements for the Mandatory<br />

Agency-Wide Quality System. Available at http://www.epa.gov/irmpoli8/policies/<br />

21050.pdf. Accessed December 3, 2012.<br />

US Environmental Protection Agency. 2002. Guidelines for Ensuring and Maximizing the Quality,<br />

Objectivity, Utility, and Integrity of Information Disseminated by EPA. US Environmental<br />

Protection Agency, Washington, DC.<br />

US Environmental Protection Agency. 2003a. A Summary of General Assessment Factors for<br />

Evaluating the Quality of Scientific and Technical Information. Available at<br />

http://www.epa.gov/spc/pdfs/assess2.pdf. Accessed November 27, 2012.<br />

US Environmental Protection Agency. 2003b. Toxic Substances Control Act (TSCA) Confidential<br />

Business Information (CBI) Protection Manual. Available at http://www.epa.gov/oppt/<br />

pubs/tsca-cbi-protection-manual.pdf. Accessed November 27, 2012.<br />

US Environmental Protection Agency. 2004a. Assuring the Competency of Environmental<br />

Protection Agency Laboratories. Available at http://epa.gov/osa/fem/pdfs/labdirective.pdf.<br />

Accessed December 3, 2012.<br />

US Environmental Protection Agency, Office of Water. 2004b. Evaluation of Impacts to<br />

Underground Sources of Drinking Water by Hydraulic Fracturing of Coalbed Methane<br />

Reservoirs. EPA 816-R-04-003. Available at http://water.epa.gov/type/groundwater/uic/<br />

class2/hydraulicfracturing/wells_coalbedmethanestudy.cfm. Accessed November 27, 2012.<br />

US Environmental Protection Agency. 2006. Office of Research and Development Policies and<br />

Procedures Manual. US Environmental Protection Agency, Washington, D.C.<br />

US Environmental Protection Agency. 2007. U.S. EPA Reach File 1 (RF1) for the Conterminous<br />

United States. Better Assessment Science Integrating point & Non-point Sources (BASINS).<br />

Washington, DC. Available at http://water.epa.gov/scitech/datait/models/basins/<br />

b3webdwn.cfm. Accessed 6/11/2007.<br />

188


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

US Environmental Protection Agency. 2008. EPA Procedure for Quality Policy. Available at<br />

http://www.epa.gov/irmpoli8/policies/2106p01.pdf. Accessed December 3, 2012.<br />

US Environmental Protection Agency. Chesapeake Bay Phase 5.3 Community Watershed Model.<br />

Available at http://www.chesapeakebay.net/about/programs/modeling/53/. Accessed<br />

November 27, 2012.<br />

US Environmental Protection Agency. 2010b. Hydraulic Fracturing Study Consultation with Tribal<br />

Governments, August 5 and 30, 2010. Available at http://www.epa.gov/<strong>hf</strong>study/<br />

publicoutreach.html. Accessed September 5, 2012.<br />

US Environmental Protection Agency. 2010c. Integrated Climate and Land-Use Scenarios (ICLUS)<br />

version 1.3 User's Manual: ArcGIS Tools and Datasets for Modeling US Housing Density<br />

Growth. Available at http://www.epa.gov/ncea/global/iclus/. Accessed September 2010.<br />

US Environmental Protection Agency. 2010d. Summary of Public Comments, Binghamton, New<br />

York, September 13 and 15, 2010. Hydraulic Fracturing Public Informational Meeting.<br />

Available at http://www.epa.gov/<strong>hf</strong>study/publicoutreach.html. Accessed September 5,<br />

2012.<br />

US Environmental Protection Agency. 2010e. Summary of Public Comments, Canonsburg,<br />

Pennsylvania, July 22, 2010. Hydraulic Fracturing Public Informational Meeting. Available at<br />

http://www.epa.gov/<strong>hf</strong>study/publicoutreach.html. Accessed September 5, 2012.<br />

US Environmental Protection Agency. 2010f. Summary of Public Comments, Denver, Colorado, July<br />

13, 2010. Hydraulic Fracturing Public Informational Meeting. Available at<br />

http://www.epa.gov/<strong>hf</strong>study/publicoutreach.html. Accessed September 5, 2012.<br />

US Environmental Protection Agency. 2010g. Summary of Public Comments, Fort Worth, Texas, July<br />

8, 2010. Hydraulic Fracturing Public Information Meeting. Available at<br />

http://www.epa.gov/<strong>hf</strong>study/publicoutreach.html. Accessed September 5, 2012.<br />

US Environmental Protection Agency. 2010h. Technical Basis for the Lowest Concentration<br />

Minimum Reporting Level (LCMRL) Calculator. EPA 815-R-11-001. Available at<br />

http://water.epa.gov/scitech/drinkingwater/labcert/upload/LCMRLTechRpt.pdf. Accessed<br />

November 27, 2012.<br />

US Environmental Protection Agency. 2011a. Counties with Oil and Gas Production Wells<br />

Hydraulically Fractured from September 2009 through October 2010 Shapefile. Data<br />

received from nine hydraulic fracturing service companies. US Environmental Protection<br />

Agency, Washington, DC.<br />

US Environmental Protection Agency. 2011b. Data Received from Hydraulic Fracturing Service<br />

Companies. Non-confidential business information is available at<br />

http://www.regulations.gov/#!docketDetail;rpp=100;so=DESC;sb=docId;po=0;D=EPA-HQ­<br />

ORD-2010-0674. Accessed November 27, 2012.<br />

189


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

US Environmental Protection Agency. 2011c. EPA Records Schedule 501. Applied and Directed<br />

Scientific Research. Available at http://www.epa.gov/records/policy/schedule/sched/<br />

501.htm. Accessed November 27, 2012.<br />

US Environmental Protection Agency. 2011d. Location of Oil and Gas Production Wells Selected for<br />

Review Shapefile. US Environmental Protection Agency, Washington, DC.<br />

US Environmental Protection Agency. 2011e. Plan to Study the Potential Impacts of Hydraulic<br />

Fracturing on Drinking Water Resources. EPA/600/R-11/122. Available at<br />

http://www.epa.gov/<strong>hf</strong>study/. Accessed November 27, 2012.<br />

US Environmental Protection Agency. 2011f. Proceedings of the US Environmental Protection<br />

Agency Technical Workshops for the Hydraulic Fracturing Study: Water Resources<br />

Management, Arlington, Virginia. Available at http://www.epa.gov/<strong>hf</strong>study/<br />

waterworkshop.html. Accessed August 24, 2012.<br />

US Environmental Protection Agency. 2011g. Quality Assurance Project Plan: Chemical<br />

Characterization of Select Constituents Relevant to Hydraulic Fracturing. Available at<br />

http://www.epa.gov/<strong>hf</strong>study/qapps.html. Accessed November 27, 2012.<br />

US Environmental Protection Agency. 2011h. Quality Assurance Project Plan: Formation of<br />

Disinfection By-products from Hydraulic Fracturing Fluid Constituents. Available at<br />

http://www.epa.gov/<strong>hf</strong>study/qapps.html. Accessed November 27, 2012.<br />

US Environmental Protection Agency. 2011i. Quality Assurance Project Plan: Hydraulic Fracturing<br />

Retrospective Case Study, Bakken Shale, Killdeer and Dunn County, North Dakota. Available<br />

at http://www.epa.gov/<strong>hf</strong>study/bakken-qapp.pdf. Accessed November 27, 2012.<br />

US Environmental Protection Agency. 2011j. Safe Drinking Water Information System. US<br />

Environmental Protection Agency, Washington, DC.<br />

US Environmental Protection Agency. 2011k. Sampling Data for Flowback and Produced Water<br />

Provided to EPA by Nine Oil and Gas Well Operators (Non-Confidential Business<br />

Information). Available at http://www.regulations.gov/<br />

#!docketDetail;rpp=100;so=DESC;sb=docId;po=0;D=EPA-HQ-ORD-2010-0674. Accessed<br />

November 27, 2012.<br />

Estimation Programs Interface Suite for Microsoft® Windows (EPI Suite) version 4.10. 2012a. US<br />

Environmental Protection Agency, Washington DC.<br />

US Environmental Protection Agency. 2012b. Exposure Assessment Tools and Models: Estimation<br />

Program Interface (EPI) Suite. Available at http://www.epa.gov/oppt/exposure/pubs/<br />

episuite.htm. Accessed November 21, 2012.<br />

US Environmental Protection Agency. 2012c. Interim EJSCREEN Common User Guidelines. Draft for<br />

Environmental Justice Committee Presentation. US Environmental Protection Agency,<br />

Washington, DC.<br />

190


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

US Environmental Protection Agency. Key Documents About Mid-Atlantic Oil and Gas Extraction.<br />

Available at http://www.epa.gov/region3/marcellus_shale/. Accessed December 5, 2012.<br />

US Environmental Protection Agency. 2012e. Peer Review Handbook. 3rd Edition. Available at<br />

http://www.epa.gov/peerreview/pdfs/peer_review_handbook_2012.pdf. Accessed<br />

December 3, 2012.<br />

US Environmental Protection Agency. 2012f. Quality Assurance Project Plan: Data and Literature<br />

Evaluation for the EPA's Study of the Potential Impacts of Hydraulic Fracturing (HF) on<br />

Drinking Water Resources. Available at http://www.epa.gov/<strong>hf</strong>study/qapps.html. Accessed<br />

December 4, 2012.<br />

US Environmental Protection Agency. 2012g. Quality Assurance Project Plan: Analysis of Data<br />

Extracted from FracFocus. Available at http://www.epa.gov/<strong>hf</strong>study/qapps.html. Accessed<br />

December 4, 2012.<br />

US Environmental Protection Agency. 2012h. Quality Assurance Project Plan: Analysis of Data<br />

Received from Nine Hydraulic Fracturing (HF) Service Companies. Available at<br />

http://www.epa.gov/<strong>hf</strong>study/qapps.html. Accessed December 4, 2012.<br />

US Environmental Protection Agency. 2012i. Quality Assurance Project Plan: Chemical Information<br />

Quality Review and Physicochemical Property Calculations for Hydraulic Fracturing<br />

Chemical Lists. Available at http://www.epa.gov/<strong>hf</strong>study/qapps.html. Accessed November<br />

30, 2012.<br />

US Environmental Protection Agency. 2012j. Quality Assurance Project Plan: Evaluation of Existing<br />

Production Well File Contents. Available at http://www.epa.gov/<strong>hf</strong>study/qapps.html.<br />

Accessed January 4, 2012.<br />

US Environmental Protection Agency. 2012k. Quality Assurance Project Plan: Health and Toxicity<br />

Theme Hydraulic Fracturing Study. Available at http://www.epa.gov/<strong>hf</strong>study/qapps.html.<br />

Accessed November 30, 2012.<br />

US Environmental Protection Agency. 2012l. Quality Assurance Project Plan: Hydraulic Fracturing<br />

(HF) Surface Spills Data Analysis. Available at http://www.epa.gov/<strong>hf</strong>study/qapps.html.<br />

Accessed December 4, 2012.<br />

US Environmental Protection Agency. 2012m. Quality Assurance Project Plan: Hydraulic Fracturing<br />

Retrospective Case Study, Bradford-Susquehanna Counties, Pennsylvania. Available at<br />

http://www.epa.gov/<strong>hf</strong>study/pdfs/bradford-review-casestudy.pdf. Accessed November 27,<br />

2012.<br />

US Environmental Protection Agency. 2012n. Quality Assurance Project Plan: Hydraulic Fracturing<br />

Retrospective Case Study, Marcellus Shale, Washington County, Pennsylvania. Available at<br />

http://www.epa.gov/<strong>hf</strong>study/washingtonco-qapp.pdf. Accessed November 27, 2012.<br />

191


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

US Environmental Protection Agency. 2012o. Quality Assurance Project Plan: Hydraulic Fracturing<br />

Retrospective Case Study, Raton Basin, Colorado. Available at http://www.epa.gov/<br />

<strong>hf</strong>study/qapps.html. Accessed November 27, 2012.<br />

US Environmental Protection Agency. 2012p. Quality Assurance Project Plan: Hydraulic Fracturing<br />

Retrospective Case Study, Wise, Texas. Available at http://www.epa.gov/<strong>hf</strong>study/barnettqapp.pdf.<br />

Accessed November 27, 2012.<br />

US Environmental Protection Agency. 2012q. Quality Assurance Project Plan: Hydraulic Fracturing<br />

Wastewater Source Apportionment. Available at http://www.epa.gov/<strong>hf</strong>study/qapps.html.<br />

Accessed November 27, 2012.<br />

US Environmental Protection Agency. 2012r. Quality Assurance Project Plan: Inter-Laboratory<br />

Verification and Validation of Diethylene Glycol, Triethylene Glycol, Tetraethylene Glycol, 2­<br />

Butoxyethanol and 2-Methoxyethanol in Ground and Surface Waters by Liquid<br />

Chromatography/Tandem Mass Spectrometry. Available at http://www.epa.gov/<strong>hf</strong>study/<br />

qapps.html. Accessed November 27, 2012.<br />

US Environmental Protection Agency. 2012s. Quality Assurance Project Plan: Surface Water<br />

Transport of Hydraulic Fracturing-Derived Waste Water. Available at http://www.epa.gov/<br />

<strong>hf</strong>study/qapps.html. Accessed November 27, 2012.<br />

US Environmental Protection Agency. 2012t. Quality Management Plan: Plan to Study the Impacts<br />

of Hydraulic Fracturing on Drinking Water Resources. Available at http://www.epa.gov/<br />

<strong>hf</strong>study/qapps.html. Accessed December 3, 2012.<br />

US Environmental Protection Agency. 2012u. Request for Information to Inform Hydraulic<br />

Fracturing Research Related to Drinking Water Resources. Federal Register 77:218 p.<br />

67361.<br />

US Environmental Protection Agency. 2012v. Science Advisory Board Staff Office Request for<br />

Nominations of Experts for the SAB Hydraulic Fracturing Advisory Panel. Federal Register<br />

77:162. p. 50505.<br />

US Environmental Protection Agency. 2012w. Unconventional Gas Drilling Projections for the US:<br />

2011-2040 AEO2012 Reference Case. National Energy Modeling System. US Environmental<br />

Protection Agency, Washington, DC.<br />

US Fish and Wildlife Service. 2008. Programmatic Biological Opinion for Water Depletions<br />

Associated with Bureau of Land Management’s Fluid Mineral Program within the Upper<br />

Colorado River Basin in Colorado. Memorandum. US Fish and Wildlife Service, Grand<br />

Junction, Colorado.<br />

US Government Accountability Office. 2012. Information on the Quantity, Quality, and Management<br />

of Water Produced during Oil and Gas Production. Energy-Water Nexus GAO-12-156.<br />

Available at http://www.gao.gov/products/GAO-12-156. Accessed December 12, 2012.<br />

192


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

US Office of Management and Budget. 2004. Final Information Quality Bulletin for Peer Review.<br />

Available at http://www.whitehouse.gov/sites/default/files/omb/assets/omb/<br />

memoranda/fy2005/m05-03.pdf. Accessed August 23, 2012.<br />

US Census Bureau. 2010a. Bradford County, Pennsylvania. State and County QuickFacts. Available at<br />

http://quickfacts.census.gov/qfd/states/42/42015.html. Accessed November 27, 2012.<br />

US Census Bureau. 2010b. Dunn County, North Dakota. State and County QuickFacts. Available at<br />

http://quickfacts.census.gov/qfd/states/38/38025.html. Accessed November 27, 2012.<br />

US Census Bureau. 2010c. Huerfano County, Colorado. State and County QuickFacts. Available at<br />

http://quickfacts.census.gov/qfd/states/08/08055.html. Accessed November 27, 2012.<br />

US Census Bureau. 2010d. Las Animas County, Colorado. State and County QuickFacts. Available at<br />

http://quickfacts.census.gov/qfd/states/08/08071.html. Accessed November 27, 2012.<br />

US Census Bureau. 2010e. Washington County, Pennsylvania. State and County Quick Facts. US<br />

Census Bureau, Available at http://quickfacts.census.gov/qfd/states/42/42125.html.<br />

Accessed December 5, 2012.<br />

US Census Bureau. 2010f. Wise County, Texas. State and County QuickFacts. Available at<br />

http://quickfacts.census.gov/qfd/states/48/48497.html. Accessed November 27, 2012.<br />

US Census Bureau. 2012a. Counties (and Equivalent) Shapefile. 2010 TIGER/Line Shapefiles. US<br />

Census Bureau, Washington, DC. Available at http://www.census.gov/cgibin/geo/shapefiles2010/main.<br />

Accessed December 4, 2012.<br />

US Census Bureau. 2012b. Places Shapefile. 2010 TIGER/Line Shapefiles. US Census Bureau,<br />

Washington, DC. Available at http://www.census.gov/cgi-bin/geo/shapefiles2010/main.<br />

Accessed December 5, 2012.<br />

US Census Bureau. 2012c. States (and Equivalent) Shapefile. 2010 TIGER/Line Shapefiles. US<br />

Census Bureau, Washington, DC. Available at http://www.census.gov/cgibin/geo/shapefiles2010/main.<br />

Accessed December 4, 2012.<br />

US Department of Agriculture. 2012. Information on Hydrologic Units and the Watershed Boundary<br />

Dataset. Available at http://www.ncgc.nrcs.usda.gov/wps/portal/nrcs/detail/national/<br />

technical/nra/dma/cid=nrcs143_0216_16. Accessed November 27, 2012.<br />

US Geological Survey. 2003. Petroleum Systems and Geologic Assessment of Oil and Gas in the<br />

Uinta-Piceance Province, Utah and Colorado. Digital Data Series DDS-69-B. Available at<br />

http://pubs.usgs.gov/dds/dds-069/dds-069-b/chapters.html. Accessed November 30,<br />

2012.<br />

US Geological Survey. 2011a. Allegheny River at Franklin, Pennsylvania (03025500). Surface-Water<br />

Monthly Statistics for the Nation. May 2006-September 2011. Available at<br />

193


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

http://waterdata.usgs.gov/nwis/inventory/site_no=03025500&agency_cd=USGS&amp.<br />

Accessed October 10, 2012.<br />

US Geological Survey. 2011b. Blacklick Creek at Josephine, Pennsylvania (03042000). Surface-<br />

Water Monthly Statistics for the Nation. May 2006-September 2011. Available at<br />

http://waterdata.usgs.gov/nwis/inventory/site_no=03042000. Accessed November 27,<br />

2012.<br />

US Geological Survey. 2011c. Information Relevant to the U.S. Geological Survey Assessment of the<br />

Middle Devonian Shale of the Appalacian Basin Province. Open-File Report 2011-1298. US<br />

Geological Survey. 22 p. Available at http://pubs.usgs.gov/of/2011/1298/. Accessed<br />

US Geological Survey. 2012a. Allegheny River at Franklin, Pennsylvania (03025500). Current<br />

Conditions for Selected Sites (Provisional). Available at http://nwis.waterdata.usgs.gov/<br />

nwis/uvsite_no=03025500. Accessed November 27, 2012.<br />

US Geological Survey. 2012b. Blacklick Creek at Josephine, Pennsylvania (03042000). Current<br />

Conditions for Selected Sites (Provisional). US Geological Survey, Available at<br />

http://nwis.waterdata.usgs.gov/nwis/uvsite_no=03042000. Accessed November 27, 2012.<br />

US House of Representatives. 2009. Appropriations Committee Report for the Department of the<br />

Interior, Environment, and Related Agencies Appropriations Bill, HR 2996. Available at<br />

http://www.gpo.gov/fdsys/pkg/CRPT-111hrpt180/pdf/CRPT-111hrpt180.pdf. Accessed<br />

December 6, 2012.<br />

US House of Representatives 2011. Chemicals Used in Hydraulic Fracturing. Available at<br />

http://democrats.energycommerce.house.gov/sites/default/files/documents/Hydraulic%2<br />

0Fracturing%20Report%204.18.11.pdf. Accessed November 27, 2012.<br />

Van Ginkel, C. G. 1996. Complete degradation of xenobiotic surfactants by consortia of aerobic<br />

microorganisms. Biodegradation 7 (2): 151-164.<br />

Van Schie, P. M. and Young L.Y. 2000. Biodegradation of phenol: mechanisms and applications.<br />

Bioremediation Journal 4 (1): 1-18.<br />

Wallis, S. 2007. The numerical solution of the advection-dispersion equation: a review of some basic<br />

principles. Acta Geophysica 55 (1): 85-94.<br />

Warner, N. R., Jackson, R. B., Darrah, T. H., Osborn, S. G., Down, A., Zhao, K., White, A. and Vengosh, A.<br />

2012. Geochemical evidence for possible natural migration of marcellus formation brine to<br />

shallow aquifers in Pennsylvania. Proceedings of the National Academy of Sciences of the<br />

United States of America 109 (30): 11961-11966.<br />

Watts, K. R. 2006a. Hydrostratigraphic Framework of the Raton, Vermejo, and Trinidad Aquifers in<br />

the Raton Basin, Las Animas County, Colorado. Scientific Investigations Report 2006-5129.<br />

US Geological Survey 37 p. Available at http://pubs.usgs.gov/sir/2006/5129/pdf/SIR06­<br />

5129_508.pdf. Accessed December 3, 2012.<br />

194


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Watts, K. R. 2006b. A Preliminary Evaluation of Vertical Separation between Production Intervals of<br />

Coalbed-Methane Wells and Water-Supply Wells in the Raton Basin, Huerfano and Las<br />

Animas Counties, Colorado, 1999-2004. Scientific Investigations Report 2006-5109. US<br />

Geological Survey. 15 p. Available at http://pubs.usgs.gov/sir/2006/5109/pdf/<br />

sir2006_5109.pdf. Accessed December 3, 2012.<br />

Westat. 2011. Quality Assurance Project Plan v. 1.1 for Hydraulic Fracturing. Westat for US<br />

Environmental Protection Agency. Available at http://www.epa.gov/<strong>hf</strong>study/qapps.html.<br />

Accessed December 11, 2012.<br />

Williams, D. R., Felbinger, J. K. and Squillace, P. J. 1993. Water Resources and the Hydrologic Effects<br />

of Coal Mining in Washington County, Pennsylvania. Open-File Report 89-620. US Geological<br />

Survey, Lemoyne, Pennsylvania.<br />

Williams, J. E., Taylor, L. E. and Low, D. J. 1998. Hydrogeology and Groundwater Quality of the<br />

Glaciated Valleys of Bradford, Tioga, and Potter Counties, Pennsylvania. Water Resources<br />

Report 68. US Geological Survey and Pennsylvania Geological Survey. 98 p. Available at<br />

http://eidmarcellus.org/wp-content/uploads/2011/06/USGS-Bradford-Report.pdf.<br />

Accessed December 5, 2012.<br />

Ziemkiewicz, P. 2011. Wastewater from Gas Development: Chemical Signatures in the Monongahela<br />

River Basin. Proceedings of the US Environmental Protection Agency Technical Workshops<br />

for the Hydraulic Fracturing Study: Water Resources Management, Arlington, Virginia.<br />

Available at http://epa.gov/<strong>hf</strong>study/waterworkshop.html. Accessed November 30, 2012.<br />

195


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Appendix A: Chemicals Identified in<br />

Hydraulic Fracturing Fluids and<br />

Wastewater<br />

This appendix contains tables of chemicals reported to be used in hydraulic fracturing fluids and<br />

chemicals detected in flowback and produced water. Sources of information include federal and<br />

state government documents, industry-provided data, and other reliable sources based on the<br />

availability of clear scientific methodology and verifiable original sources; the full list of<br />

information sources is available in Section A.1. The EPA at this time has not made any judgment<br />

about the extent of exposure to these chemicals when used in hydraulic fracturing fluids or found in<br />

hydraulic fracturing wastewater, or their potential impacts on drinking water resources.<br />

The tables in this appendix include information provided by nine hydraulic fracturing service<br />

companies (see Section 3.3), nine oil and gas operators (Section 3.4), and FracFocus (Section 3.5).<br />

Over 150 entries in Tables A-1 and A-2 were provided by the service companies, and roughly 60<br />

entries were provided by FracFocus; these entries were not included in easily obtained public<br />

sources. The nine oil and gas operators provided data on chemicals and properties of flowback and<br />

produced water; the chemicals and properties are listed in Tables A-3 and A-4.<br />

Much of the information provided in response to the EPA’s September 2010 information request to<br />

the nine hydraulic fracturing service companies was claimed as confidential business information<br />

(CBI) under the Toxic Substances Control Act. In many cases, the service companies have agreed to<br />

publicly release chemical names and Chemical Abstract Services Registration Numbers (CASRNs) in<br />

Table A-1. However, 82 chemicals with known chemical names and CASRNs continue to be claimed<br />

as CBI, and are not included in this appendix. In some instances, generic chemical names have been<br />

provided for these chemicals in Table A-2.<br />

In order to standardize chemical names, chemical name and structure annotation quality control<br />

methods have been applied to chemicals with CASRNs. 88 These methods ensure correct chemical<br />

names and CASRNs and include combining duplicates where appropriate.<br />

The EPA is creating a Distributed Structure-Searchable Toxicity (DSSTox) 89 chemical inventory for<br />

chemicals reported to be used in hydraulic fracturing fluids and/or detected in flowback and<br />

produced water. The hydraulic fracturing DSSTox chemical inventory will contain CASRNs,<br />

chemical names and synonyms, and structure data files (where available). The structure data files<br />

can be used with existing computer software to calculate physicochemical properties, as described<br />

in Chapter 6.<br />

88 Additional information on this process can be found at http://www.epa.gov/ncct/dsstox/<br />

ChemicalInfQAProcedures.html.<br />

89 The DSSTox website provides a public forum for publishing downloadable, structure-searchable, standardized chemical<br />

structure files associated with chemical inventories or toxicity datasets of environmental relevance. For more<br />

information, see http://www.epa.gov/ncct/dsstox/.<br />

196


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table A-1 lists chemicals reported to be used in hydraulic fracturing fluids between 2005 and 2011.<br />

This table lists chemicals with their associated CASRNs. Structure data files are expected to be in<br />

the hydraulic fracturing DSSTox chemical inventory for some chemicals on Table A-1; these<br />

chemicals are marked with a “” in the “IUPAC Name and Structure” column.<br />

Table A-1. List of CASRNs and names of chemicals reportedly used in hydraulic fracturing fluids. Chemical<br />

structures and IUPAC names have been identified for the chemicals with an “” in the “IUPAC Name and Structure”<br />

column. A few chemicals have structures, but no assigned CASRNs; these chemicals have “NA” entered in the<br />

CASRN column.<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

120086-58-0<br />

(13Z)-N,N-bis(2-hydroxyethyl)-N-methyldocos-13-en-1­<br />

aminium chloride<br />

1<br />

123-73-9 (E)-Crotonaldehyde 1, 4<br />

2235-43-0<br />

[Nitrilotris(methylene)]tris-phosphonic acid pentasodium<br />

salt<br />

1<br />

65322-65-8 1-(1-Naphthylmethyl)quinolinium chloride 1<br />

68155-37-3<br />

1-(Alkyl* amino)-3-aminopropane *(42%C12, 26%C18,<br />

15%C14, 8%C16, 5%C10, 4%C8)<br />

68909-18-2 1-(Phenylmethyl)pyridinium Et Me derivs., chlorides <br />

8<br />

526-73-8 1,2,3-Trimethylbenzene 1, 4<br />

1, 2, 3, 4,<br />

6, 8<br />

95-63-6 1,2,4-Trimethylbenzene 1, 2, 3, 4, 5<br />

2634-33-5 1,2-Benzisothiazolin-3-one 1, 3, 4<br />

35691-65-7 1,2-Dibromo-2,4-dicyanobutane 1, 4<br />

95-47-6 1,2-Dimethylbenzene 4<br />

138879-94-4<br />

1,2-Ethanediaminium, N, N'-bis[2-[bis(2­<br />

hydroxyethyl)methylammonio]ethyl]-N,N'bis(2­<br />

hydroxyethyl)-N,N'-dimethyl-,tetrachloride<br />

1, 4<br />

57-55-6 1,2-Propanediol 1, 2, 3, 4, 8<br />

75-56-9 1,2-Propylene oxide 1, 4<br />

4719-04-4 1,3,5-Triazine-1,3,5(2H,4H,6H)-triethanol 1, 4<br />

108-67-8 1,3,5-Trimethylbenzene 1, 4<br />

123-91-1 1,4-Dioxane 2, 3, 4<br />

9051-89-2<br />

1,4-Dioxane-2,5-dione, 3,6-dimethyl-, (3R,6R)-, polymer<br />

with (3S,6S)-3,6-dimethyl-1,4-dioxane-2,5-dione and<br />

(3R,6S)-rel-3,6-dimethyl-1,4-dioxane-2,5-dione<br />

1, 4, 8<br />

124-09-4 1,6-Hexanediamine 1, 2<br />

6055-52-3 1,6-Hexanediamine dihydrochloride 1<br />

20324-33-8<br />

1-[2-(2-Methoxy-1-methylethoxy)-1-methylethoxy]-2­<br />

propanol<br />

4<br />

78-96-6 1-Amino-2-propanol 8<br />

Table continued on next page<br />

197


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

15619-48-4 1-Benzylquinolinium chloride 1, 3, 4 <br />

71-36-3 1-Butanol 1, 2, 3, 4, 7 <br />

112-30-1 1-Decanol 1, 4 <br />

2687-96-9 1-Dodecyl-2-pyrrolidinone 1, 4 <br />

3452-07-1 1-Eicosene 3<br />

629-73-2 1-Hexadecene 3<br />

111-27-3 1-Hexanol 1, 4, 8 <br />

68909-68-7<br />

68442-97-7<br />

1-Hexanol, 2-ethyl-, manuf. of, by products from, distn.<br />

residues<br />

1H-Imidazole-1-ethanamine, 4,5-dihydro-, 2-nortall-oil<br />

alkyl derivs.<br />

107-98-2 1-Methoxy-2-propanol 1, 2, 3, 4 <br />

2190-04-7 1-Octadecanamine, acetate (1:1) 8<br />

124-28-7 1-Octadecanamine, N,N-dimethyl­ 1, 3, 4 <br />

112-88-9 1-Octadecene 3<br />

111-87-5 1-Octanol 1, 4 <br />

71-41-0 1-Pentanol 8<br />

61789-39-7<br />

61789-40-0<br />

68139-30-0<br />

149879-98-1<br />

1-Propanaminium, 3-amino-N-(carboxymethyl)-N,Ndimethyl-,<br />

N-coco acyl derivs., chlorides, sodium salts<br />

1-Propanaminium, 3-amino-N-(carboxymethyl)-N,Ndimethyl-,<br />

N-coco acyl derivs., inner salts<br />

1-Propanaminium, N-(3-aminopropyl)-2-hydroxy-N,Ndimethyl-3-sulfo-,<br />

N-coco acyl derivs., inner salts<br />

1-Propanaminium, N-(carboxymethyl)-N,N-dimethyl-3­<br />

[[(13Z)-1-oxo-13-docosen-1-yl]amino]-,<br />

5284-66-2 1-Propanesulfonic acid 3<br />

<br />

4<br />

2, 4 <br />

1<br />

1, 2, 3, 4 <br />

1, 3, 4 <br />

71-23-8 1-Propanol 1, 2, 4, 5 <br />

23519-77-9 1-Propanol, zirconium(4+) salt 1, 4, 8 <br />

115-07-1 1-Propene 2<br />

1120-36-1 1-Tetradecene 3<br />

112-70-9 1-Tridecanol 1, 4 <br />

112-42-5 1-Undecanol 2<br />

112-34-5 2-(2-Butoxyethoxy)ethanol 2, 4 <br />

111-90-0 2-(2-Ethoxyethoxy)ethanol 1, 4 <br />

112-15-2 2-(2-Ethoxyethoxy)ethyl acetate 1, 4 <br />

102-81-8 2-(Dibutylamino)ethanol 1, 4 <br />

1, 3 <br />

Table continued on next page<br />

198


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

34375-28-5 2-(Hydroxymethylamino)ethanol 1, 4 <br />

21564-17-0 2-(Thiocyanomethylthio)benzothiazole 2<br />

27776-21-2<br />

10213-78-2 2,2'-(Octadecylimino)diethanol 1<br />

929-59-9 2,2'-[Ethane-1,2-diylbis(oxy)]diethanamine 1, 4 <br />

9003-11-6 2,2'-[propane-1,2-diylbis(oxy)]diethanol 1, 3, 4, 8 <br />

25085-99-8<br />

2,2'-(Azobis(1-methylethylidene))bis(4,5-dihydro-1Himidazole)dihydrochloride<br />

2,2'-[propane-2,2-diylbis(4,1­<br />

phenyleneoxymethylene)]dioxirane<br />

10222-01-2 2,2-Dibromo-3-nitrilopropionamide <br />

73003-80-2 2,2-Dibromopropanediamide 3<br />

24634-61-5 2,4-Hexadienoic acid, potassium salt, (2E,4E)­ 3<br />

915-67-3<br />

2,7-Naphthalenedisulfonic acid, 3-hydroxy-4-[2-(4-sulfo­<br />

1-naphthalenyl) diazenyl] -, sodium salt (1:3)<br />

<br />

<br />

<br />

3<br />

1, 4, 8 <br />

1, 2, 3, 4, <br />

6, 7, 8 <br />

9002-93-1 2-[4-(1,1,3,3-tetramethylbutyl)phenoxy]ethanol 1, 3, 4 <br />

NA<br />

2-Acrylamide - 2-propanesulfonic acid and N,Ndimethylacrylamide<br />

copolymer<br />

NA 2-acrylamido -2-methylpropanesulfonic acid copolymer 2<br />

15214-89-8 2-Acrylamido-2-methyl-1-propanesulfonic acid 1, 3 <br />

124-68-5 2-Amino-2-methylpropan-1-ol 8<br />

2002-24-6 2-Aminoethanol hydrochloride 4, 8 <br />

52-51-7 2-Bromo-2-nitropropane-1,3-diol 1, 2, 3, 4, 6 <br />

1113-55-9 2-Bromo-3-nitrilopropionamide 1, 2, 3, 4, 5 <br />

96-29-7 2-Butanone oxime 1<br />

143106-84-7<br />

68442-77-3<br />

2-Butanone, 4-[[[(1R,4aS,10aR)-1,2,3,4,4a,9,10,10a­<br />

octahydro-1,4a-dimethyl-7-(1-methylethyl)-1­<br />

phenanthrenyl]methyl](3-oxo-3-phenylpropyl)amino]-,<br />

hydrochloride (1:1)<br />

2-Butenediamide, (2E)-, N,N'-bis[2-(4,5-dihydro-2-nortalloil<br />

alkyl-1H-imidazol-1-yl)ethyl] derivs.<br />

111-76-2 2-Butoxyethanol <br />

110-80-5 2-Ethoxyethanol 6<br />

<br />

<br />

4<br />

2<br />

1, 4 <br />

3, 8 <br />

1, 2, 3, 4, <br />

6, 7, 8 <br />

104-76-7 2-Ethyl-1-hexanol 1, 2, 3, 4, 5 <br />

645-62-5 2-Ethyl-2-hexenal 2<br />

5444-75-7 2-Ethylhexyl benzoate 4<br />

Table continued on next page<br />

199


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

818-61-1 2-Hydroxyethyl acrylate 1, 4<br />

13427-63-9 2-Hydroxyethylammonium hydrogen sulphite 1<br />

60-24-2 2-Mercaptoethanol 1, 4<br />

109-86-4 2-Methoxyethanol 4<br />

78-83-1 2-Methyl-1-propanol 1, 2, 4<br />

107-41-5 2-Methyl-2,4-pentanediol 1, 2, 4<br />

2682-20-4 2-Methyl-3(2H)-isothiazolone 1, 2, 4<br />

115-19-5 2-Methyl-3-butyn-2-ol 3<br />

78-78-4 2-Methylbutane 2<br />

62763-89-7 2-Methylquinoline hydrochloride 3<br />

37971-36-1 2-Phosphono-1,2,4-butanetricarboxylic acid 1, 4<br />

93858-78-7<br />

2-Phosphonobutane-1,2,4-tricarboxylic acid, potassium<br />

salt (1:x)<br />

1<br />

555-31-7 2-Propanol, aluminum salt 1<br />

26062-79-3<br />

2-Propen-1-aminium, N,N-dimethyl-N-2-propenyl-,<br />

chloride, homopolymer<br />

<br />

3<br />

13533-05-6 2-Propenoic acid, 2-(2-hydroxyethoxy)ethyl ester 4<br />

113221-69-5<br />

111560-38-4<br />

2-Propenoic acid, ethyl ester, polymer with ethenyl<br />

acetate and 2,5-furandione, hydrolyzed<br />

2-Propenoic acid, ethyl ester, polymer with ethenyl<br />

acetate and 2,5-furandione, hydrolyzed, sodium salt<br />

4, 8<br />

8<br />

9003-04-7 2-Propenoic acid, homopolymer, sodium salt 1, 2, 3, 4<br />

9003-06-9 2-Propenoic acid, polymer with 2-propenamide 4, 8<br />

25987-30-8<br />

37350-42-8<br />

151006-66-5<br />

2-Propenoic acid, polymer with 2-propenamide, sodium<br />

salt<br />

2-Propenoic acid, sodium salt (1:1), polymer with sodium<br />

2-methyl-2-((1-oxo-2-propen-1-yl)amino)-1­<br />

propanesulfonate (1:1)<br />

2-Propenoic acid, telomer with sodium 4­<br />

ethenylbenzenesulfonate (1:1), sodium 2-methyl-2-[(1­<br />

oxo-2-propen-1-yl)amino]-1-propanesulfonate (1:1) and<br />

sodium sulfite (1:1), sodium salt<br />

1<br />

71050-62-9 2-Propenoic, polymer with sodium phosphinate 3, 4<br />

75673-43-7 3,4,4-Trimethyloxazolidine 8<br />

51229-78-8<br />

3,5,7-Triazatricyclo(3.3.1.1(superscript 3,7))decane, 1-(3­<br />

chloro-2-propenyl)-, chloride, (Z)­<br />

3, 4, 8<br />

4<br />

3<br />

5392-40-5 3,7-Dimethyl-2,6-octadienal 3<br />

Table continued on next page<br />

200


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

104-55-2 3-Phenylprop-2-enal 1, 2, 3, 4, 7<br />

12068-08-5 4-(Dodecan-6-yl)benzenesulfonic acid – morpholine (1:1) 1, 4<br />

51200-87-4 4,4-Dimethyloxazolidine 8<br />

5877-42-9 4-Ethyloct-1-yn-3-ol 1, 2, 3, 4<br />

121-33-5 4-Hydroxy-3-methoxybenzaldehyde 3<br />

122-91-8 4-Methoxybenzyl formate 3<br />

150-76-5 4-Methoxyphenol 4<br />

108-11-2 4-Methyl-2-pentanol 1, 4<br />

108-10-1 4-Methyl-2-pentanone 5<br />

104-40-5 4-Nonylphenol 8<br />

26172-55-4 5-Chloro-2-methyl-3(2H)-isothiazolone 1, 2, 4<br />

106-22-9 6-Octen-1-ol, 3,7-dimethyl­ 3<br />

75-07-0 Acetaldehyde 1, 4<br />

64-19-7 Acetic acid <br />

1, 2, 3, 4,<br />

5, 6, 7, 8<br />

25213-24-5 Acetic acid ethenyl ester, polymer with ethenol 1, 4<br />

90438-79-2 Acetic acid, C6-8-branched alkyl esters 4<br />

68442-62-6<br />

Acetic acid, hydroxy-, reaction products with<br />

triethanolamine<br />

3<br />

5421-46-5 Acetic acid, mercapto-, monoammonium salt 2, 8<br />

108-24-7 Acetic anhydride 1, 2, 3, 4, 7<br />

67-64-1 Acetone 1, 3, 4, 6<br />

7327-60-8 Acetonitrile, 2,2',2''-nitrilotris­ 1, 4<br />

98-86-2 Acetophenone 1<br />

77-89-4 Acetyltriethyl citrate 1, 4<br />

107-02-8 Acrolein 2<br />

79-06-1 Acrylamide 1, 2, 3, 4<br />

25085-02-3 Acrylamide/ sodium acrylate copolymer 1, 2, 3, 4, 8<br />

38193-60-1<br />

Acrylamide-sodium-2-acrylamido-2-methlypropane<br />

sulfonate copolymer<br />

1, 2, 3, 4<br />

79-10-7 Acrylic acid 2, 4<br />

110224-99-2<br />

Acrylic acid, with sodium-2-acrylamido-2-methyl-1­<br />

propanesulfonate and sodium phosphinate<br />

8<br />

67254-71-1 Alcohols, C10-12, ethoxylated 3<br />

68526-86-3 Alcohols, C11-14-iso-, C13-rich 3<br />

Table continued on next page<br />

201


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

CASRN Chemical Name<br />

228414-35-5 Alcohols, C11-14-iso-, C13-rich, butoxylated ethoxylated<br />

78330-21-9 Alcohols, C11-14-iso-, C13-rich, ethoxylated<br />

126950-60-5 Alcohols, C12-14-secondary<br />

84133-50-6 Alcohols, C12-14-secondary, ethoxylated<br />

78330-19-5 Alcohols, C7-9-iso-, C8-rich, ethoxylated<br />

68603-25-8 Alcohols, C8-10, ethoxylated propoxylated<br />

78330-20-8 Alcohols, C9-11-iso-, C10-rich, ethoxylated<br />

93924-07-3 Alkanes, C10-14<br />

90622-52-9 Alkanes, C10-16-branched and linear<br />

68551-19-9 Alkanes, C12-14-iso­<br />

68551-20-2 Alkanes, C13-16-iso­<br />

64743-02-8 Alkenes, C>10 .alpha.­<br />

68411-00-7 Alkenes, C>8<br />

68607-07-8<br />

Alkenes, C24-25 alpha-, polymers with maleic anhydride,<br />

docosyl esters<br />

71011-24-0 Alkyl quaternary ammonium with bentonite<br />

85409-23-0<br />

Alkyl* dimethyl ethylbenzyl ammonium chloride<br />

*(50%C12, 30%C14, 17%C16, 3%C18)<br />

42615-29-2 Alkylbenzenesulfonate, linear<br />

1302-62-1 Almandite and pyrope garnet<br />

60828-78-6<br />

alpha-[3.5-dimethyl-1-(2-methylpropyl)hexyl]-omegahydroxy-poly(oxy-1,2-ethandiyl)<br />

9000-90-2 alpha-Amylase<br />

98-55-5 Alpha-Terpineol<br />

1302-42-7 Aluminate (AlO 1- 2 ), sodium<br />

7429-90-5 Aluminum<br />

12042-68-1 Aluminum calcium oxide (Al 2 CaO 4 )<br />

7446-70-0 Aluminum chloride<br />

1327-41-9 Aluminum chloride, basic<br />

1344-28-1 Aluminum oxide<br />

12068-56-3 Aluminum oxide silicate<br />

12141-46-7 Aluminum silicate<br />

10043-01-3 Aluminum sulfate<br />

68155-07-7 Amides, C8-18 and C18-unsatd., N,N-bis(hydroxyethyl)<br />

68140-01-2 Amides, coco, N-[3-(dimethylamino)propyl]<br />

IUPAC<br />

Name and Reference<br />

Structure<br />

1<br />

3, 4, 8<br />

1, 3, 4 <br />

3, 4, 8 <br />

2, 4, 8<br />

3<br />

1, 2, 4, 8<br />

1<br />

4<br />

2, 4, 8 <br />

1, 4<br />

1, 3, 4, 8 <br />

1<br />

8<br />

4<br />

8<br />

1, 4, 6 <br />

1, 4 <br />

3<br />

4<br />

3<br />

2, 4<br />

1, 4, 6 <br />

2<br />

1, 4 <br />

3, 4<br />

1, 2, 4 <br />

1, 2, 4<br />

1, 2, 4 <br />

1, 4 <br />

3<br />

1, 4 <br />

Table continued on next page<br />

202


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

70851-07-9<br />

Amides, coco, N-[3-(dimethylamino)propyl], alkylation<br />

products with chloroacetic acid, sodium salts<br />

68155-09-9 Amides, coco, N-[3-(dimethylamino)propyl], N-oxides 1, 3, 4 <br />

68876-82-4 Amides, from C16-22 fatty acids and diethylenetriamine 3<br />

68155-20-4 Amides, tall-oil fatty, N,N-bis(hydroxyethyl) 3, 4 <br />

68647-77-8 Amides, tallow, N-[3-(dimethylamino)propyl],N-oxides 1, 4 <br />

68155-39-5 Amines, C14-18; C16-18-unsaturated, alkyl, ethoxylated 1<br />

68037-94-5 Amines, C8-18 and C18-unsatd. alkyl 5<br />

61788-46-3 Amines, coco alkyl 4<br />

61790-57-6 Amines, coco alkyl, acetates 1, 4 <br />

61788-93-0 Amines, coco alkyldimethyl 8<br />

61790-59-8 Amines, hydrogenated tallow alkyl, acetates 4<br />

68966-36-9<br />

68603-67-8<br />

Amines, polyethylenepoly-, ethoxylated,<br />

phosphonomethylated<br />

Amines, polyethylenepoly-, reaction products with benzyl<br />

chloride<br />

61790-33-8 Amines, tallow alkyl 8<br />

61791-26-2 Amines, tallow alkyl, ethoxylated 1, 3 <br />

68551-33-7 Amines, tallow alkyl, ethoxylated, acetates (salts) 1, 3, 4 <br />

68308-48-5 Amines, tallow alkyl, ethoxylated, phosphates 4<br />

6419-19-8 Aminotrimethylene phosphonic acid 1, 4, 8 <br />

7664-41-7 Ammonia 1, 2, 3, 4, 7 <br />

32612-48-9 Ammonium (lauryloxypolyethoxy)ethyl sulfate 4<br />

631-61-8 Ammonium acetate 1, 3, 4, 5, 8 <br />

10604-69-0 Ammonium acrylate 8<br />

26100-47-0 Ammonium acrylate-acrylamide polymer 2, 4, 8 <br />

7803-63-6 Ammonium bisulfate 2<br />

10192-30-0 Ammonium bisulfite 1, 2, 3, 4, 7 <br />

12125-02-9 Ammonium chloride <br />

7632-50-0 Ammonium citrate (1:1) 3<br />

3012-65-5 Ammonium citrate (2:1) 8<br />

2235-54-3 Ammonium dodecyl sulfate 1<br />

<br />

1, 4 <br />

1, 4 <br />

1<br />

1, 2, 3, 4, <br />

5, 6, 8 <br />

12125-01-8 Ammonium fluoride 1, 4 <br />

1066-33-7 Ammonium hydrogen carbonate 1, 4 <br />

Table continued on next page<br />

203


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

1341-49-7 Ammonium hydrogen difluoride 1, 3, 4, 7 <br />

13446-12-3 Ammonium hydrogen phosphonate 4<br />

1336-21-6 Ammonium hydroxide 1, 3, 4 <br />

8061-53-8 Ammonium ligninsulfonate 2<br />

6484-52-2 Ammonium nitrate 1, 2, 3 <br />

7722-76-1 Ammonium phosphate 1, 4 <br />

7783-20-2 Ammonium sulfate 1, 2, 3, 4, 6 <br />

99439-28-8 Amorphous silica 1, 7 <br />

104-46-1 Anethole 3<br />

62-53-3 Aniline 2, 4 <br />

1314-60-9 Antimony pentoxide 1, 4 <br />

10025-91-9 Antimony trichloride 1, 4 <br />

1309-64-4 Antimony trioxide 8<br />

7440-38-2 Arsenic 4<br />

68131-74-8 Ashes, residues 4<br />

68201-32-1 Asphalt, sulfonated, sodium salt 2<br />

12174-11-7 Attapulgite 2, 3 <br />

31974-35-3 Aziridine, polymer with 2-methyloxirane 4, 8 <br />

7727-43-7 Barium sulfate 1, 2, 4 <br />

1318-16-7 Bauxite 1, 2, 4 <br />

1302-78-9 Bentonite 1, 2, 4, 6 <br />

121888-68-4<br />

Bentonite, benzyl(hydrogenated tallow alkyl) <br />

dimethylammonium stearate complex<br />

80-08-0 Benzamine, 4,4'-sulfonylbis­ 1, 4 <br />

71-43-2 Benzene 1, 3, 4 <br />

98-82-8 Benzene, (1-methylethyl)­ 1, 2, 3, 4 <br />

119345-03-8 Benzene, 1,1'-oxybis-, tetrapropylene derivs., sulfonated 8<br />

119345-04-9<br />

Benzene, 1,1'-oxybis-, tetrapropylene derivs., sulfonated,<br />

sodium salts<br />

611-14-3 Benzene, 1-ethyl-2-methyl­ 4<br />

68648-87-3 Benzene, C10-16-alkyl derivs. 1<br />

9003-55-8 Benzene, ethenyl-, polymer with 1,3-butadiene 2, 4 <br />

74153-51-8<br />

Benzenemethanaminium, N,N-dimethyl-N-(2-((1-oxo-2­<br />

propen-1-yl)oxy)ethyl)-, chloride (1:1), polymer with 2­<br />

propenamide<br />

98-11-3 Benzenesulfonic acid 2<br />

<br />

3, 4 <br />

3, 4, 8 <br />

3<br />

Table continued on next page<br />

204


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

37953-05-2 Benzenesulfonic acid, (1-methylethyl)-, 4<br />

37475-88-0 Benzenesulfonic acid, (1-methylethyl)-, ammonium salt 3, 4 <br />

28348-53-0 Benzenesulfonic acid, (1-methylethyl)-, sodium salt 8<br />

68584-22-5 Benzenesulfonic acid, C10-16-alkyl derivs. 1, 4 <br />

255043-08-4<br />

68584-27-0<br />

90218-35-2<br />

Benzenesulfonic acid, C10-16-alkyl derivs., compds. with<br />

cyclohexylamine<br />

Benzenesulfonic acid, C10-16-alkyl derivs., potassium<br />

salts<br />

Benzenesulfonic acid, dodecyl-, branched, compds. with<br />

2-propanamine<br />

26264-06-2 Benzenesulfonic acid, dodecyl-, calcium salt 4<br />

68648-81-7<br />

Benzenesulfonic acid, mono-C10-16 alkyl derivs.,<br />

compds. with 2-propanamine<br />

<br />

<br />

<br />

<br />

1<br />

1, 4, 8 <br />

65-85-0 Benzoic acid 1, 4, 7 <br />

100-44-7 Benzyl chloride 1, 2, 4, 8 <br />

139-07-1 Benzyldimethyldodecylammonium chloride 2, 8 <br />

122-18-9 Benzylhexadecyldimethylammonium chloride 8<br />

68425-61-6<br />

Bis(1-methylethyl)naphthalenesulfonic acid,<br />

cyclohexylamine salt<br />

111-44-4 Bis(2-chloroethyl) ether 8<br />

80-05-7 Bisphenol A 4<br />

65996-69-2 Blast furnace slag 2, 3 <br />

1303-96-4 Borax 1, 2, 3, 4, 6 <br />

10043-35-3 Boric acid <br />

<br />

4<br />

1, 4 <br />

1<br />

1, 2, 3, 4, <br />

6, 7 <br />

1303-86-2 Boric oxide 1, 2, 3, 4 <br />

11128-29-3 Boron potassium oxide 1<br />

1330-43-4 Boron sodium oxide 1, 2, 4 <br />

12179-04-3 Boron sodium oxide pentahydrate 8<br />

106-97-8 Butane 2, 5 <br />

2373-38-8<br />

Butanedioic acid, sulfo-, 1,4-bis(1,3-dimethylbutyl) ester,<br />

sodium salt<br />

2673-22-5 Butanedioic acid, sulfo-, 1,4-ditridecyl ester, sodium salt 4<br />

2426-08-6 Butyl glycidyl ether 1, 4 <br />

138-22-7 Butyl lactate 1, 4 <br />

3734-67-6 C.I. Acid red 1 4<br />

<br />

1<br />

Table continued on next page<br />

205


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

6625-46-3 C.I. Acid violet 12, disodium salt 4<br />

6410-41-9 C.I. Pigment Red 5 4<br />

4477-79-6 C.I. Solvent Red 26 4<br />

70592-80-2 C10-16-Alkyldimethylamines oxides 4<br />

68002-97-1 C10-C16 ethoxylated alcohol 1, 2, 3, 4, 8 <br />

68131-40-8 C11-15-Secondary alcohols ethoxylated 1, 2, 8 <br />

73138-27-9 C12-14 tert-alkyl ethoxylated amines 3<br />

66402-68-4 Calcined bauxite 2, 8 <br />

12042-78-3 Calcium aluminate 2<br />

7789-41-5 Calcium bromide 4<br />

10043-52-4 Calcium chloride 1, 2, 3, 4, 7 <br />

10035-04-8 Calcium dichloride dihydrate 1, 4 <br />

7789-75-5 Calcium fluoride 1, 4 <br />

1305-62-0 Calcium hydroxide 1, 2, 3, 4 <br />

7778-54-3 Calcium hypochlorite 1, 2, 4 <br />

58398-71-3 Calcium magnesium hydroxide oxide 4<br />

1305-78-8 Calcium oxide 1, 2, 4, 7 <br />

1305-79-9 Calcium peroxide 1, 3, 4, 8 <br />

7778-18-9 Calcium sulfate 1, 2, 4 <br />

10101-41-4 Calcium sulfate dihydrate 2<br />

76-22-2 Camphor 3<br />

1333-86-4 Carbon black 1, 2, 4 <br />

124-38-9 Carbon dioxide 1, 3, 4, 6 <br />

471-34-1 Carbonic acid calcium salt (1:1) 1, 2, 4 <br />

584-08-7 Carbonic acid, dipotassium salt 1, 2, 3, 4, 8 <br />

39346-76-4 Carboxymethyl guar gum, sodium salt 1, 2, 4 <br />

61791-12-6 Castor oil, ethoxylated 1, 3 <br />

8000-27-9 Cedarwood oil 3<br />

9005-81-6 Cellophane 1, 4 <br />

9012-54-8 Cellulase 1, 2, 3, 4, 5 <br />

9004-34-6 Cellulose 1, 2, 3, 4 <br />

9004-32-4 Cellulose, carboxymethyl ether, sodium salt 2, 3, 4 <br />

16887-00-6 Chloride 4, 8 <br />

7782-50-5 Chlorine 2<br />

Table continued on next page<br />

206


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

10049-04-4 Chlorine dioxide 1, 2, 3, 4, 8 <br />

78-73-9 Choline bicarbonate 3, 8 <br />

67-48-1 Choline chloride 1, 3, 4, 7, 8 <br />

16065-83-1 Chromium (III), insoluble salts 2, 6 <br />

18540-29-9 Chromium (VI) 6<br />

39430-51-8 Chromium acetate, basic 2<br />

1066-30-4 Chromium(III) acetate 1, 2 <br />

77-92-9 Citric acid 1, 2, 3, 4, 7 <br />

8000-29-1 Citronella oil 3<br />

94266-47-4 Citrus extract 1, 3, 4, 8 <br />

50815-10-6 Coal, granular 1, 2, 4 <br />

71-48-7 Cobalt(II) acetate 1, 4 <br />

68424-94-2 Coco-betaine 3<br />

68603-42-9 Coconut oil acid/Diethanolamine condensate (2:1) 1<br />

61789-18-2 Coconut trimethylammonium chloride 1, 8 <br />

7440-50-8 Copper 1, 4 <br />

7758-98-7 Copper sulfate 1, 4, 8 <br />

7758-89-6 Copper(I) chloride 1, 4 <br />

7681-65-4 Copper(I) iodide 1, 2, 4, 6 <br />

7447-39-4 Copper(II) chloride 1, 3, 4 <br />

68525-86-0 Corn flour 4<br />

11138-66-2 Corn sugar gum 1, 2, 4 <br />

1302-74-5 Corundum (Aluminum oxide) 4, 8 <br />

68308-87-2 Cottonseed, flour 2, 4 <br />

91-64-5 Coumarin 3<br />

14464-46-1 Cristobalite 1, 2, 4 <br />

15468-32-3 Crystalline silica, tridymite 1, 2, 4 <br />

10125-13-0 Cupric chloride dihydrate 1, 4, 7 <br />

110-82-7 Cyclohexane 1, 7 <br />

108-94-1 Cyclohexanone 1, 4 <br />

18472-87-2 D&C Red 28 4<br />

533-74-4 Dazomet <br />

1, 2, 3, 4, <br />

7, 8 <br />

1120-24-7 Decyldimethylamine 3, 4 <br />

7789-20-0 Deuterium oxide 8<br />

Table continued on next page<br />

207


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

50-70-4 D-Glucitol 1, 3, 4 <br />

526-95-4 D-Gluconic acid 1, 4 <br />

3149-68-6 D-Glucopyranoside, methyl 2<br />

50-99-7 D-Glucose 1, 4 <br />

117-81-7 Di(2-ethylhexyl) phthalate 1, 4 <br />

7727-54-0 Diammonium peroxydisulfate <br />

68855-54-9 Diatomaceous earth 2, 4 <br />

1, 2, 3, 4, <br />

6, 7, 8 <br />

91053-39-3 Diatomaceous earth, calcined 1, 2, 4 <br />

3252-43-5 Dibromoacetonitrile 1, 2, 3, 4, 8 <br />

10034-77-2 Dicalcium silicate 1, 2, 4 <br />

7173-51-5 Didecyldimethylammonium chloride 1, 2, 4, 8 <br />

111-42-2 Diethanolamine 1, 2, 3, 4, 6 <br />

25340-17-4 Diethylbenzene 1, 3, 4 <br />

111-46-6 Diethylene glycol 1, 2, 3, 4, 7 <br />

111-77-3 Diethylene glycol monomethyl ether 1, 2, 4 <br />

111-40-0 Diethylenetriamine 1, 2, 4, 5 <br />

68647-57-4 Diethylenetriamine reaction product with fatty acid dimers 2<br />

38640-62-9 Diisopropylnaphthalene 3, 4 <br />

627-93-0 Dimethyl adipate 8<br />

1119-40-0 Dimethyl glutarate 1, 4 <br />

63148-62-9 Dimethyl polysiloxane 1, 2, 4 <br />

106-65-0 Dimethyl succinate 8<br />

108-01-0 Dimethylaminoethanol 2, 4 <br />

7398-69-8 Dimethyldiallylammonium chloride 3, 4 <br />

101-84-8 Diphenyl oxide 3<br />

7758-11-4 Dipotassium monohydrogen phosphate 5<br />

25265-71-8 Dipropylene glycol 1, 3, 4 <br />

31291-60-8 Di-sec-butylphenol 1<br />

28519-02-0<br />

Disodium <br />

dodecyl(sulphonatophenoxy)benzenesulphonate<br />

38011-25-5 Disodium ethylenediaminediacetate 1, 4 <br />

6381-92-6 Disodium ethylenediaminetetraacetate dihydrate 1<br />

12008-41-2 Disodium octaborate 4, 8 <br />

12280-03-4 Disodium octaborate tetrahydrate 1, 4 <br />

<br />

1<br />

Table continued on next page<br />

208


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

68477-31-6<br />

68333-25-5<br />

Distillates, petroleum, catalytic reformer fractionator<br />

residue, low-boiling<br />

Distillates, petroleum, hydrodesulfurized light catalytic<br />

cracked<br />

64742-80-9 Distillates, petroleum, hydrodesulfurized middle 1<br />

64742-52-5 Distillates, petroleum, hydrotreated heavy naphthenic 1, 2, 3, 4<br />

64742-54-7 Distillates, petroleum, hydrotreated heavy paraffinic 1, 2, 4<br />

64742-47-8 Distillates, petroleum, hydrotreated light<br />

1, 4<br />

1<br />

1, 2, 3, 4,<br />

5, 7, 8<br />

64742-53-6 Distillates, petroleum, hydrotreated light naphthenic 1, 2, 8<br />

64742-55-8 Distillates, petroleum, hydrotreated light paraffinic 8<br />

64742-46-7 Distillates, petroleum, hydrotreated middle 1, 2, 3, 4, 8<br />

64741-59-9 Distillates, petroleum, light catalytic cracked 1, 4<br />

64741-77-1 Distillates, petroleum, light hydrocracked 3<br />

64742-65-0 Distillates, petroleum, solvent-dewaxed heavy paraffinic 1<br />

64741-96-4 Distillates, petroleum, solvent-refined heavy naphthenic 1, 4<br />

64742-91-2 Distillates, petroleum, steam-cracked 1, 4<br />

64741-44-2 Distillates, petroleum, straight-run middle 1, 2, 4<br />

64741-86-2 Distillates, petroleum, sweetened middle 1, 4<br />

71011-04-6 Ditallow alkyl ethoxylated amines 3<br />

10326-41-7 D-Lactic acid 1, 4<br />

5989-27-5 D-Limonene <br />

577-11-7 Docusate sodium 1<br />

112-40-3 Dodecane 8<br />

123-01-3 Dodecylbenzene 3, 4<br />

1, 3, 4, 5,<br />

7, 8<br />

27176-87-0 Dodecylbenzene sulfonic acid 2, 3, 4, 8<br />

26836-07-7 Dodecylbenzenesulfonic acid, monoethanolamine salt 1, 4<br />

12276-01-6 EDTA, copper salt 1, 5, 6<br />

37288-54-3 Endo-1,4-.beta.-mannanase. 3, 8<br />

106-89-8 Epichlorohydrin 1, 4, 8<br />

44992-01-0<br />

69418-26-4<br />

Ethanaminium, N,N,N-trimethyl-2-[(1-oxo-2­<br />

propenyl)oxy]-, chloride<br />

Ethanaminium, N,N,N-trimethyl-2-[(1-oxo-2­<br />

propenyl)oxy]-, chloride, polymer with 2-propenamide<br />

3<br />

1, 3, 4<br />

Table continued on next page<br />

209


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

CASRN<br />

Chemical Name<br />

Ethanaminium, N,N,N-trimethyl-2[(2-methyl-1-oxo-2­<br />

26006-22-4 propen-1-yl0oxy]-, methyl sulfate 91:1), polymer with 2­<br />

propenamide<br />

27103-90-8<br />

74-84-0 Ethane<br />

64-17-5 Ethanol<br />

68171-29-9<br />

Ethanaminium, N,N,N-trimethyl-2-[(2-methyl-1-oxo-2­<br />

propenyl)oxy]-, methyl sulfate, homopolymer<br />

Ethanol, 2,2',2''-nitrilotris-, tris(dihydrogen phosphate)<br />

(ester), sodium salt<br />

61791-47-7 Ethanol, 2,2'-iminobis-, N-coco alkyl derivs., N-oxides<br />

61791-44-4 Ethanol, 2,2'-iminobis-, N-tallow alkyl derivs.<br />

68909-77-3<br />

68877-16-7<br />

Ethanol, 2,2'-oxybis-, reaction products with ammonia,<br />

morpholine derivs. residues<br />

Ethanol, 2,2-oxybis-, reaction products with ammonia,<br />

morpholine derivs. residues, acetates (salts)<br />

Ethanol, 2,2-oxybis-, reaction products with ammonia,<br />

102424-23-7 morpholine derivs. residues, reaction products with sulfur<br />

dioxide<br />

25446-78-0<br />

Ethanol, 2-[2-[2-(tridecyloxy)ethoxy]ethoxy]-, hydrogen<br />

sulfate, sodium salt<br />

34411-42-2 Ethanol, 2-amino-, polymer with formaldehyde<br />

68649-44-5<br />

141-43-5 Ethanolamine<br />

Ethanol, 2-amino-, reaction products with ammonia, byproducts<br />

from, phosphonomethylated<br />

66455-15-0 Ethoxylated C10-14 alcohols<br />

66455-14-9 Ethoxylated C12-13 alcohols<br />

68439-50-9 Ethoxylated C12-14 alcohols<br />

68131-39-5 Ethoxylated C12-15 alcohols<br />

68551-12-2 Ethoxylated C12-16 alcohols<br />

68951-67-7 Ethoxylated C14-15 alcohols<br />

68439-45-2 Ethoxylated C6-12 alcohols<br />

68439-46-3 Ethoxylated C9-11 alcohols<br />

9002-92-0 Ethoxylated dodecyl alcohol<br />

61790-82-7 Ethoxylated hydrogenated tallow alkylamines<br />

68439-51-0 Ethoxylated propoxylated C12-14 alcohols<br />

52624-57-4 Ethoxylated, propoxylated trimethylolpropane<br />

141-78-6 Ethyl acetate<br />

IUPAC<br />

Name and<br />

Structure<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Reference<br />

1, 4<br />

8<br />

2, 5<br />

1, 2, 3, 4,<br />

5, 6, 8 <br />

4<br />

1<br />

1<br />

4, 8<br />

4<br />

4<br />

1, 4<br />

4<br />

4<br />

1, 2, 3, 4, 6 <br />

3<br />

4<br />

2, 3, 4, 8 <br />

3, 4<br />

3, 4, 8<br />

3, 4, 8 <br />

3, 4, 8 <br />

3, 4<br />

4<br />

4<br />

1, 3, 4, 8 <br />

3<br />

1, 4, 7 <br />

Table continued on next page<br />

210


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

141-97-9 Ethyl acetoacetate 1, 4 <br />

93-89-0 Ethyl benzoate 3<br />

97-64-3 Ethyl lactate 3<br />

118-61-6 Ethyl salicylate 3<br />

100-41-4 Ethylbenzene 1, 2, 3, 4, 7 <br />

9004-57-3 Ethylcellulose 2<br />

107-21-1 Ethylene glycol <br />

1, 2, 3, 4, <br />

6, 7, 8 <br />

75-21-8 Ethylene oxide 1, 2, 3, 4 <br />

107-15-3 Ethylenediamine 2, 4 <br />

60-00-4 Ethylenediaminetetraacetic acid 1, 2, 4 <br />

64-02-8 Ethylenediaminetetraacetic acid tetrasodium salt 1, 2, 3, 4 <br />

67989-88-2<br />

Ethylenediaminetetraacetic acid, diammonium copper<br />

salt<br />

139-33-3 Ethylenediaminetetraacetic acid, disodium salt 1, 3, 4, 8 <br />

74-86-2 Ethyne 7<br />

68604-35-3<br />

Fatty acids, C 8-18 and C18-unsaturated compounds<br />

with diethanolamine<br />

70321-73-2 Fatty acids, C14-18 and C16-18-unsatd., distn. residues 2<br />

61788-89-4 Fatty acids, C18-unsatd., dimers 2<br />

61791-29-5 Fatty acids, coco, ethoxylated 3<br />

61791-08-0<br />

Fatty acids, coco, reaction products with ethanolamine,<br />

ethoxylated<br />

61790-90-7 Fatty acids, tall oil, hexa esters with sorbitol, ethoxylated 1, 4 <br />

68188-40-9<br />

Fatty acids, tall oil, reaction products with acetophenone,<br />

formaldehyde and thiourea<br />

61790-12-3 Fatty acids, tall-oil 1, 2, 3, 4 <br />

61790-69-0<br />

Fatty acids, tall-oil, reaction products with<br />

diethylenetriamine<br />

8052-48-0 Fatty acids, tallow, sodium salts 1, 3 <br />

68153-72-0<br />

Fatty acids, vegetable-oil, reaction products with<br />

diethylenetriamine<br />

3844-45-9 FD&C Blue no. 1 1, 4 <br />

7705-08-0 Ferric chloride 1, 3, 4 <br />

10028-22-5 Ferric sulfate 1, 4 <br />

17375-41-6 Ferrous sulfate monohydrate 2<br />

<br />

4<br />

3<br />

3<br />

3<br />

1, 4 <br />

3<br />

Table continued on next page<br />

211


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

65997-17-3 Fiberglass 2, 3, 4 <br />

50-00-0 Formaldehyde 1, 2, 3, 4 <br />

NA Formaldehyde amine 8<br />

29316-47-0<br />

63428-92-2<br />

28906-96-9<br />

30704-64-4<br />

Formaldehyde polymer with 4,1,1-(dimethylethyl)phenol<br />

and methyloxirane<br />

Formaldehyde polymer with methyl oxirane, 4­<br />

nonylphenol and oxirane<br />

Formaldehyde, polymer with 2-(chloromethyl)oxirane and<br />

4,4'-(1-methylethylidene)bis[phenol]<br />

Formaldehyde, polymer with 4-(1,1-dimethylethyl)phenol,<br />

2-methyloxirane and oxirane<br />

<br />

<br />

<br />

<br />

3<br />

4, 8 <br />

1, 4 <br />

30846-35-6 Formaldehyde, polymer with 4-nonylphenol and oxirane 1, 4 <br />

35297-54-2 Formaldehyde, polymer with ammonia and phenol 1, 4 <br />

25085-75-0 Formaldehyde, polymer with bisphenol A 4<br />

70750-07-1<br />

Formaldehyde, polymer with N1-(2-aminoethyl)-1,2­<br />

ethanediamine, benzylated<br />

55845-06-2 Formaldehyde, polymer with nonylphenol and oxirane 4<br />

153795-76-7<br />

Formaldehyde, polymers with branched 4-nonylphenol,<br />

ethylene oxide and propylene oxide<br />

<br />

<br />

1, 2, 4, 8 <br />

75-12-7 Formamide 1, 2, 3, 4 <br />

64-18-6 Formic acid <br />

8<br />

1, 3 <br />

1, 2, 3, 4, <br />

6, 7 <br />

590-29-4 Formic acid, potassium salt 1, 3, 4 <br />

68476-30-2 Fuel oil, no. 2 1, 2 <br />

68334-30-5 Fuels, diesel 2<br />

68476-34-6 Fuels, diesel, no. 2 2, 4, 8 <br />

8031-18-3 Fuller's earth 2<br />

110-17-8 Fumaric acid 1, 2, 3, 4, 6 <br />

98-01-1 Furfural 1, 4 <br />

98-00-0 Furfuryl alcohol 1, 4 <br />

64741-43-1 Gas oils, petroleum, straight-run 1, 4 <br />

9000-70-8 Gelatin 1, 4 <br />

12002-43-6 Gilsonite 1, 2, 4 <br />

133-42-6 Gluconic acid 7<br />

111-30-8 Glutaraldehyde 1, 2, 3, 4, 7 <br />

56-81-5 Glycerin, natural 1, 2, 3, 4, 5 <br />

Table continued on next page<br />

212


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

135-37-5<br />

Glycine, N-(carboxymethyl)-N-(2-hydroxyethyl)-,<br />

disodium salt<br />

150-25-4 Glycine, N,N-bis(2-hydroxyethyl)­ 1, 4 <br />

5064-31-3 Glycine, N,N-bis(carboxymethyl)-, trisodium salt 1, 2, 3, 4 <br />

139-89-9<br />

Glycine, N-[2-[bis(carboxymethyl)amino]ethyl]-N-(2­<br />

hydroxyethyl)-, trisodium salt<br />

79-14-1 Glycolic acid 1, 3, 4 <br />

2836-32-0 Glycolic acid sodium salt 1, 3, 4 <br />

107-22-2 Glyoxal 1, 2, 4 <br />

298-12-4 Glyoxylic acid 1<br />

9000-30-0 Guar gum<br />

68130-15-4<br />

Guar gum, carboxymethyl 2-hydroxypropyl ether, sodium<br />

salt<br />

13397-24-5 Gypsum 2, 4 <br />

67891-79-6 Heavy aromatic distillate 1, 4 <br />

<br />

<br />

1<br />

1<br />

1, 2, 3, 4, <br />

7, 8 <br />

1, 2, 3, 4, 7 <br />

1317-60-8 Hematite 1, 2, 4 <br />

9025-56-3 Hemicellulase enzyme concentrate 3, 4 <br />

142-82-5 Heptane 1, 2 <br />

68526-88-5 Heptene, hydroformylation products, high-boiling 1, 4 <br />

57-09-0 Hexadecyltrimethylammonium bromide 1<br />

110-54-3 Hexane 5<br />

124-04-9 Hexanedioic acid 1, 2, 4, 6 <br />

1415-93-6 Humic acids, commercial grade 2<br />

68956-56-9 Hydrocarbons, terpene processing by-products 1, 3, 4 <br />

7647-01-0 Hydrochloric acid <br />

1, 2, 3, 4, <br />

5, 6, 7, 8 <br />

7664-39-3 Hydrogen fluoride 1, 2, 4 <br />

7722-84-1 Hydrogen peroxide 1, 3, 4 <br />

7783-06-4 Hydrogen sulfide 1, 2 <br />

9004-62-0 Hydroxyethylcellulose 1, 2, 3, 4 <br />

4719-04-4 Hydroxylamine hydrochloride 1, 3, 4 <br />

10039-54-0 Hydroxylamine sulfate (2:1) 4<br />

9004-64-2 Hydroxypropyl cellulose 2, 4 <br />

39421-75-5 Hydroxypropyl guar gum<br />

1, 3, 4, 5, <br />

6, 8 <br />

Table continued on next page<br />

213


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

120-72-9 Indole 2<br />

430439-54-6 Inulin, carboxymethyl ether, sodium salt 1, 4 <br />

12030-49-8 Iridium oxide 8<br />

7439-89-6 Iron 2, 4 <br />

1317-61-9 Iron oxide (Fe3O4) 4<br />

1332-37-2 Iron(II) oxide 1, 4 <br />

7720-78-7 Iron(II) sulfate 2<br />

7782-63-0 Iron(II) sulfate heptahydrate 1, 2, 3, 4 <br />

1309-37-1 Iron(III) oxide 1, 2, 4 <br />

89-65-6 Isoascorbic acid 1, 3, 4 <br />

75-28-5 Isobutane 2<br />

26952-21-6 Isooctanol 1, 4, 5 <br />

123-51-3 Isopentyl alcohol 1, 4 <br />

67-63-0 Isopropanol <br />

1, 2, 3, 4, <br />

6, 7 <br />

42504-46-1 Isopropanolamine dodecylbenzenesulfonate 1, 3, 4 <br />

75-31-0 Isopropylamine 1, 4 <br />

68909-80-8<br />

Isoquinoline, reaction products with benzyl chloride and<br />

quinoline<br />

35674-56-7 Isoquinolinium, 2-(phenylmethyl)-, chloride 3<br />

9043-30-5 Isotridecanol, ethoxylated 1, 3, 4, 8 <br />

1332-58-7 Kaolin 1, 2, 4 <br />

8008-20-6 Kerosine (petroleum) 1, 2, 3, 4, 8 <br />

64742-81-0 Kerosine, petroleum, hydrodesulfurized 1, 2, 4 <br />

61790-53-2 Kieselguhr 1, 2, 4 <br />

1302-76-7 Kyanite 1, 2, 4 <br />

50-21-5 Lactic acid 1, 4, 8 <br />

63-42-3 Lactose 3<br />

13197-76-7 Lauryl hydroxysultaine 1<br />

8022-15-9 Lavandula hybrida abrial herb oil 3<br />

4511-42-6 L-Dilactide 1, 4 <br />

7439-92-1 Lead 1, 4 <br />

8002-43-5 Lecithin 4<br />

129521-66-0 Lignite 2<br />

8062-15-5 Lignosulfuric acid 2<br />

<br />

3<br />

Table continued on next page<br />

214


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

1317-65-3 Limestone 1, 2, 3, 4 <br />

8001-26-1 Linseed oil 8<br />

79-33-4 L-Lactic acid 1, 4, 8 <br />

546-93-0 Magnesium carbonate (1:1) 1, 3, 4 <br />

7786-30-3 Magnesium chloride 1, 2, 4 <br />

7791-18-6 Magnesium chloride hexahydrate 4<br />

1309-42-8 Magnesium hydroxide 1, 4 <br />

19086-72-7 Magnesium iron silicate 1, 4 <br />

10377-60-3 Magnesium nitrate 1, 2, 4 <br />

1309-48-4 Magnesium oxide 1, 2, 3, 4 <br />

14452-57-4 Magnesium peroxide 1, 4 <br />

12057-74-8 Magnesium phosphide 1<br />

1343-88-0 Magnesium silicate 1, 4 <br />

26099-09-2 Maleic acid homopolymer 8<br />

25988-97-0<br />

Methanamine-N-methyl polymer with chloromethyl<br />

oxirane<br />

74-82-8 Methane 2, 5 <br />

67-56-1 Methanol <br />

<br />

4<br />

1, 2, 3, 4, <br />

5, 6, 7, 8 <br />

100-97-0 Methenamine 1, 2, 4 <br />

625-45-6 Methoxyacetic acid 8<br />

9004-67-5 Methyl cellulose 8<br />

119-36-8 Methyl salicylate 1, 2, 3, 4, 7 <br />

78-94-4 Methyl vinyl ketone 1, 4 <br />

108-87-2 Methylcyclohexane 1<br />

6317-18-6 Methylene bis(thiocyanate) 2<br />

66204-44-2 Methylenebis(5-methyloxazolidine) 2<br />

68891-11-2<br />

Methyloxirane polymer with oxirane, mono (nonylphenol)<br />

ether, branched<br />

12001-26-2 Mica 1, 2, 4, 6 <br />

8012-95-1 Mineral oil - includes paraffin oil 4, 8 <br />

64475-85-0 Mineral spirits 2<br />

26038-87-9 Monoethanolamine borate (1:x) 1, 4 <br />

1318-93-0 Montmorillonite 2<br />

110-91-8 Morpholine 1, 2, 4 <br />

<br />

3<br />

Table continued on next page<br />

215


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

78-21-7 Morpholinium, 4-ethyl-4-hexadecyl-, ethyl sulfate 8<br />

1302-93-8 Mullite 1,2, 4, 8 <br />

46830-22-2<br />

54076-97-0<br />

19277-88-4<br />

N-(2-Acryloyloxyethyl)-N-benzyl-N,N-dimethylammonium<br />

chloride<br />

N,N,N-Trimethyl-2[1-oxo-2-propenyl]oxy ethanaminimum<br />

chloride, homopolymer<br />

N,N,N-Trimethyl-3-((1-oxooctadecyl)amino)-1­<br />

propanaminium methyl sulfate<br />

112-03-8 N,N,N-Trimethyloctadecan-1-aminium chloride 1, 3, 4 <br />

109-46-6 N,N'-Dibutylthiourea 1, 4 <br />

2605-79-0 N,N-Dimethyldecylamine oxide 1, 3, 4 <br />

68-12-2 N,N-Dimethylformamide 1, 2, 4, 5, 8 <br />

593-81-7 N,N-Dimethylmethanamine hydrochloride 1, 4, 5, 7 <br />

1184-78-7 N,N-Dimethyl-methanamine-N-oxide 3<br />

1613-17-8 N,N-Dimethyloctadecylamine hydrochloride 1, 4 <br />

110-26-9 N,N'-Methylenebisacrylamide 1, 4 <br />

64741-68-0 Naphtha, petroleum, heavy catalytic reformed 1, 2, 3, 4 <br />

64742-48-9 Naphtha, petroleum, hydrotreated heavy 1, 2, 3, 4, 8 <br />

91-20-3 Naphthalene <br />

93-18-5 Naphthalene, 2-ethoxy­ 3<br />

<br />

<br />

<br />

3<br />

3<br />

1<br />

1, 2, 3, 4, <br />

5, 7 <br />

28757-00-8 Naphthalenesulfonic acid, bis(1-methylethyl)- 1, 3, 4 <br />

99811-86-6<br />

Naphthalenesulphonic acid, bis (1-methylethyl)-methyl<br />

derivatives<br />

68410-62-8 Naphthenic acid ethoxylate 4<br />

7786-81-4 Nickel sulfate 2<br />

10101-97-0 Nickel(II) sulfate hexahydrate 1, 4 <br />

61790-29-2 Nitriles, tallow, hydrogenated 4<br />

4862-18-4 Nitrilotriacetamide 1, 4, 7 <br />

139-13-9 Nitrilotriacetic acid 1, 4 <br />

18662-53-8 Nitrilotriacetic acid trisodium monohydrate 1, 4 <br />

7727-37-9 Nitrogen 1, 2, 3, 4, 6 <br />

872-50-4 N-Methyl-2-pyrrolidone 1, 4 <br />

105-59-9 N-Methyldiethanolamine 2, 4, 8 <br />

109-83-1 N-Methylethanolamine 4<br />

68213-98-9 N-Methyl-N-hydroxyethyl-N-hydroxyethoxyethylamine 4<br />

<br />

1<br />

Table continued on next page<br />

216


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

13127-82-7 N-Oleyl diethanolamide 1, 4<br />

25154-52-3 Nonylphenol (mixed) 1, 4<br />

8000-48-4 Oil of eucalyptus 3<br />

8007-02-1 Oil of lemongrass 3<br />

8000-25-7 Oil of rosemary 3<br />

112-80-1 Oleic acid 2, 4<br />

1317-71-1 Olivine 4<br />

8028-48-6 Orange terpenes 4<br />

68649-29-6<br />

51838-31-4<br />

Oxirane, methyl-, polymer with oxirane, mono-C10-16­<br />

alkyl ethers, phosphates <br />

Oxiranemethanaminium, N,N,N-trimethyl-, chloride,<br />

homopolymer<br />

7782-44-7 Oxygen 4<br />

10028-15-6 Ozone 8<br />

8002-74-2 Paraffin waxes and Hydrocarbon waxes 1<br />

30525-89-4 Paraformaldehyde 2<br />

4067-16-7 Pentaethylenehexamine 4<br />

109-66-0 Pentane 2, 5<br />

628-63-7 Pentyl acetate 3<br />

540-18-1 Pentyl butyrate 3<br />

79-21-0 Peracetic acid 8<br />

93763-70-3 Perlite 4<br />

64743-01-7 Petrolatum, petroleum, oxidized 3<br />

8002-05-9 Petroleum 1, 2<br />

6742-47-8 Petroleum distillate hydrotreated light 8<br />

85-01-8 Phenanthrene 6<br />

<br />

1, 4<br />

1, 2, 3, 4,<br />

5, 8<br />

108-95-2 Phenol 1, 2, 4<br />

25068-38-6<br />

Phenol, 4,4'-(1-methylethylidene)bis-, polymer with 2­<br />

(chloromethyl)oxirane<br />

1, 2, 4<br />

9003-35-4 Phenol, polymer with formaldehyde 1, 2, 4, 7<br />

7803-51-2 Phosphine 1, 4<br />

13598-36-2 Phosphonic acid 1, 4<br />

29712-30-9 Phosphonic acid (dimethylamino(methylene)) 1<br />

129828-36-0<br />

Phosphonic acid, (((2-[(2­<br />

hydroxyethyl)(phosphonomethyl)amino)ethyl)imino]bis(m<br />

ethylene))bis-, compd. with 2-aminoethanol<br />

1<br />

Table continued on next page<br />

217


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

67953-76-8<br />

3794-83-0<br />

15827-60-8<br />

70714-66-8<br />

22042-96-2<br />

34690-00-1<br />

Phosphonic acid, (1-hydroxyethylidene)bis-, potassium<br />

salt<br />

Phosphonic acid, (1-hydroxyethylidene)bis-, tetrasodium<br />

salt<br />

Phosphonic acid, [[(phosphonomethyl)imino]bis[2,1­<br />

ethanediylnitrilobis(methylene)]]tetrakis-<br />

Phosphonic acid, [[(phosphonomethyl)imino]bis[2,1­<br />

ethanediylnitrilobis(methylene)]]tetrakis-, ammonium salt<br />

(1:x) <br />

Phosphonic acid, [[(phosphonomethyl)imino]bis[2,1­<br />

ethanediylnitrilobis(methylene)]]tetrakis-, sodium salt<br />

Phosphonic acid, [[(phosphonomethyl)imino]bis[6,1­<br />

hexanediylnitrilobis(methylene)]]tetrakis­<br />

<br />

<br />

<br />

<br />

<br />

<br />

4<br />

1, 4 <br />

1, 2, 4 <br />

3<br />

3<br />

1, 4, 8 <br />

7664-38-2 Phosphoric acid 1, 2, 4 <br />

7785-88-8 Phosphoric acid, aluminium sodium salt 1, 2 <br />

7783-28-0 Phosphoric acid, diammonium salt 2<br />

68412-60-2 Phosphoric acid, mixed decyl and Et and octyl esters 1<br />

10294-56-1 Phosphorous acid 1<br />

85-44-9 Phthalic anhydride 1, 4 <br />

8002-09-3 Pine oils 1, 2, 4 <br />

25038-54-4 Policapram (Nylon 6) 1, 4 <br />

62649-23-4 Poly (acrylamide-co-acrylic acid), partial sodium salt 3, 4 <br />

26680-10-4 Poly(lactide) 1<br />

9014-93-1<br />

9016-45-9<br />

51811-79-1<br />

68987-90-6<br />

26635-93-8<br />

9004-96-0<br />

68891-38-3<br />

Poly(oxy-1,2-ethanediyl), .alpha.-(dinonylphenyl)-. <br />

omega.-hydroxy­<br />

Poly(oxy-1,2-ethanediyl), .alpha.-(nonylphenyl)-.omega.­<br />

hydroxy­<br />

Poly(oxy-1,2-ethanediyl), .alpha.-(nonylphenyl)-.omega.­<br />

hydroxy-, phosphate<br />

Poly(oxy-1,2-ethanediyl), .alpha.-(octylphenyl)-.omega.­<br />

hydroxy-, branched <br />

Poly(oxy-1,2-ethanediyl), .alpha.,.alpha.'-[[(9Z)-9­<br />

octadecenylimino]di-2,1-ethanediyl]bis[.omega.-hydroxy­<br />

Poly(oxy-1,2-ethanediyl), .alpha.-[(9Z)-1-oxo-9­<br />

octadecenyl]-.omega.-hydroxy­<br />

Poly(oxy-1,2-ethanediyl), .alpha.-sulfo-.omega.-hydroxy-,<br />

C12-14-alkyl ethers, sodium salts<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

4<br />

1, 2, 3, 4, 8 <br />

1, 4 <br />

1, 4 <br />

1, 4 <br />

8<br />

1, 4 <br />

Table continued on next page<br />

218


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

61723-83-9<br />

68015-67-8<br />

68412-53-3<br />

Poly(oxy-1,2-ethanediyl), a-hydro-w-hydroxy-, ether with<br />

D-glucitol (2:1), tetra-(9Z)-9-octadecenoate <br />

Poly(oxy-1,2-ethanediyl), alpha-(2,3,4,5­<br />

tetramethylnonyl)-omega-hydroxy<br />

Poly(oxy-1,2-ethanediyl), alpha-(nonylphenyl)-omegahydroxy-,branched,<br />

phosphates<br />

8<br />

1<br />

1<br />

31726-34-8 Poly(oxy-1,2-ethanediyl), alpha-hexyl-omega-hydroxy 3, 8<br />

56449-46-8<br />

65545-80-4<br />

27306-78-1<br />

52286-19-8<br />

63428-86-4<br />

68037-05-8<br />

9081-17-8<br />

52286-18-7<br />

68890-88-0<br />

Poly(oxy-1,2-ethanediyl), alpha-hydro-omega-hydroxy-,<br />

(9Z)-9-octadecenoate <br />

Poly(oxy-1,2-ethanediyl), alpha-hydro-omega-hydroxy-,<br />

ether with alpha-fluoro-omega-(2­<br />

hydroxyethyl)poly(difluoromethylene) (1:1) <br />

Poly(oxy-1,2-ethanediyl), alpha-methyl-omega-(3­<br />

(1,3,3,3-tetramethyl-1-((trimethylsilyl)oxy)-1­<br />

disiloxanyl)propoxy)­<br />

Poly(oxy-1,2-ethanediyl), alpha-sulfo-omega-(decyloxy)-,<br />

ammonium salt (1:1) <br />

Poly(oxy-1,2-ethanediyl), alpha-sulfo-omega-(hexyloxy)-,<br />

ammonium salt (1:1) <br />

Poly(oxy-1,2-ethanediyl), alpha-sulfo-omega-(hexyloxy)-,<br />

C6-10-alkyl ethers, ammonium salts<br />

Poly(oxy-1,2-ethanediyl), alpha-sulfo-omega-­<br />

(nonylphenoxy)­<br />

Poly(oxy-1,2-ethanediyl), alpha-sulfo-omega-(octyloxy) -, <br />

ammonium salt (1:1) <br />

Poly(oxy-1,2-ethanediyl), alpha-sulfo-omega-hydroxy-,<br />

C10-12-alkyl ethers, ammonium salts<br />

<br />

<br />

<br />

<br />

<br />

3<br />

1<br />

1<br />

4<br />

1, 3, 4<br />

3, 4<br />

4<br />

4<br />

8<br />

24938-91-8 Poly(oxy-1,2-ethanediyl), alpha-tridecyl-omega-hydroxy­ 1, 3, 4<br />

127036-24-2<br />

68412-54-4<br />

Poly(oxy-1,2-ethanediyl), alpha-undecyl-omega-hydroxy-,<br />

branched and linear<br />

1<br />

2, 3, 4<br />

34398-01-1 Poly-(oxy-1,2-ethanediyl)-alpha-undecyl-omega-hydroxy 1, 3, 4, 8<br />

127087-87-0 Poly(oxy-1,2-ethanediyl)-nonylphenyl-hydroxy branched 1, 2, 3, 4<br />

25704-18-1 Poly(sodium-p-styrenesulfonate) 1, 4<br />

32131-17-2<br />

Poly(oxy-1,2-ethanediyl),alpha-(4-nonylphenyl)-omegahydroxy-,branched<br />

Poly[imino(1,6-dioxo-1,6-hexanediyl)imino-1,6­<br />

hexanediyl]<br />

2<br />

9003-05-8 Polyacrylamide 1, 2, 4, 6<br />

Table continued on next page<br />

219


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

NA Polyacrylate/ polyacrylamide blend 2<br />

66019-18-9 Polyacrylic acid, sodium bisulfite terminated 3<br />

25322-68-3 Polyethylene glycol <br />

9004-98-2 Polyethylene glycol (9Z)-9-octadecenyl ether 8<br />

68187-85-9 Polyethylene glycol ester with tall oil fatty acid 1<br />

1, 2, 3, 4, <br />

7, 8 <br />

9036-19-5 Polyethylene glycol mono(octylphenyl) ether 1, 2, 3, 4, 8 <br />

9004-77-7 Polyethylene glycol monobutyl ether 1, 4 <br />

68891-29-2<br />

Polyethylene glycol mono-C8-10-alkyl ether sulfate<br />

ammonium<br />

<br />

1, 3, 4 <br />

9046-01-9 Polyethylene glycol tridecyl ether phosphate 1, 3, 4 <br />

9002-98-6 Polyethyleneimine 4<br />

25618-55-7 Polyglycerol 2<br />

9005-70-3 Polyoxyethylene sorbitan trioleate 3<br />

26027-38-3 Polyoxyethylene(10)nonylphenyl ether 1, 2, 3, 4, 8 <br />

9046-10-0 Polyoxypropylenediamine 1<br />

68131-72-6<br />

Polyphosphoric acids, esters with triethanolamine, <br />

sodium salts<br />

68915-31-1 Polyphosphoric acids, sodium salts 1, 4 <br />

25322-69-4 Polypropylene glycol 1, 2, 4 <br />

68683-13-6<br />

Polypropylene glycol glycerol triether, epichlorohydrin,<br />

bisphenol A polymer<br />

9011-19-2 Polysiloxane 4<br />

9005-64-5 Polysorbate 20 8<br />

9003-20-7 Polyvinyl acetate copolymer 2<br />

9002-89-5 Polyvinyl alcohol 1, 2, 4 <br />

NA Polyvinyl alcohol/polyvinyl acetate copolymer 1<br />

9002-85-1 Polyvinylidene chloride 8<br />

65997-15-1 Portland cement 2, 4 <br />

127-08-2 Potassium acetate 1, 3, 4 <br />

1327-44-2 Potassium aluminum silicate 5<br />

29638-69-5 Potassium antimonate 1, 4 <br />

12712-38-8 Potassium borate 3<br />

20786-60-1 Potassium borate (1:x) 1, 3 <br />

6381-79-9 Potassium carbonate sesquihydrate 5<br />

1<br />

1<br />

Table continued on next page<br />

220


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

7447-40-7 Potassium chloride <br />

7778-50-9 Potassium dichromate 4<br />

1, 2, 3, 4,<br />

5, 6, 7<br />

1310-58-3 Potassium hydroxide 1, 2, 3, 4, 6<br />

7681-11-0 Potassium iodide 1, 4<br />

13709-94-9 Potassium metaborate 1, 2, 3, 4, 8<br />

143-18-0 Potassium oleate 4<br />

12136-45-7 Potassium oxide 1, 4<br />

7727-21-1 Potassium persulfate 1, 2, 4<br />

7778-80-5 Potassium sulfate 2<br />

74-98-6 Propane 2, 5<br />

2997-92-4<br />

Propanimidamide,2,2’' -aAzobis[(2-methyl-,<br />

amidinopropane) dihydrochloride<br />

1, 4<br />

34590-94-8 Propanol, 1(or 2)-(2-methoxymethylethoxy)­ 1, 2, 3, 4<br />

107-19-7 Propargyl alcohol <br />

108-32-7 Propylene carbonate 1, 4<br />

15220-87-8 Propylene pentamer 1<br />

106-42-3 p-Xylene 1, 4<br />

68391-11-7 Pyridine, alkyl derivs. 1, 4<br />

100765-57-9 Pyridinium, 1-(phenylmethyl)-, alkyl derivs., chlorides 4, 8<br />

70914-44-2<br />

Pyridinium, 1-(phenylmethyl)-, C7-8-alkyl derivs.,<br />

chlorides<br />

6<br />

289-95-2 Pyrimidine 2<br />

109-97-7 Pyrrole 2<br />

14808-60-7 Quartz <br />

308074-31-9<br />

68607-28-3<br />

68153-30-0<br />

68989-00-4<br />

Quaternary ammonium compounds (2-ethylhexyl)<br />

hydrogenated tallow alkyl)dimethyl, methyl sulfates<br />

Quaternary ammonium compounds, (oxydi-2,1­<br />

ethanediyl)bis[coco alkyldimethyl, dichlorides<br />

Quaternary ammonium compounds,<br />

benzylbis(hydrogenated tallow alkyl)methyl, salts with<br />

bentonite<br />

Quaternary ammonium compounds, benzyl-C10-16­<br />

alkyldimethyl, chlorides<br />

1, 2, 3, 4,<br />

5, 6, 7, 8<br />

1, 2, 3, 4,<br />

5, 6, 8<br />

8<br />

2, 3, 4, 8<br />

2, 5, 6<br />

1, 4<br />

Table continued on next page<br />

221


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

68424-85-1<br />

68391-01-5<br />

61789-68-2<br />

68953-58-2<br />

71011-27-3<br />

68424-95-3<br />

61789-77-3<br />

68607-29-4<br />

8030-78-2<br />

Quaternary ammonium compounds, benzyl-C12-16­<br />

alkyldimethyl, chlorides<br />

Quaternary ammonium compounds, benzyl-C12-18­<br />

alkyldimethyl, chlorides <br />

Quaternary ammonium compounds, benzylcoco<br />

alkylbis(hydroxyethyl), chlorides<br />

Quaternary ammonium compounds, bis(hydrogenated<br />

tallow alkyl)dimethyl, salts with bentonite<br />

Quaternary ammonium compounds, bis(hydrogenated<br />

tallow alkyl)dimethyl, salts with hectorite <br />

Quaternary ammonium compounds, di-C8-10­<br />

alkyldimethyl, chlorides <br />

Quaternary ammonium compounds, dicoco alkyldimethyl,<br />

chlorides<br />

Quaternary ammonium compounds, pentamethyltallow<br />

alkyltrimethylenedi-, dichlorides <br />

Quaternary ammonium compounds, trimethyltallow alkyl,<br />

chlorides<br />

<br />

1, 2, 4, 8<br />

8<br />

1, 4<br />

2, 3, 4, 8<br />

2<br />

2<br />

91-22-5 Quinoline 2, 4<br />

68514-29-4 Raffinates (petroleum) 5<br />

64741-85-1 Raffinates, petroleum, sorption process 1, 2, 4, 8<br />

64742-01-4 Residual oils, petroleum, solvent-refined 5<br />

64741-67-9 Residues, petroleum, catalytic reformer fractionator 1, 4, 8<br />

81-88-9 Rhodamine B 4<br />

8050-09-7 Rosin 1, 4<br />

12060-08-1 Scandium oxide 8<br />

63800-37-3 Sepiolite 2<br />

68611-44-9 Silane, dichlorodimethyl-, reaction products with silica 2<br />

7631-86-9 Silica 1, 2, 3, 4, 8<br />

112926-00-8 Silica gel, cryst. -free 3, 4<br />

112945-52-5 Silica, amorphous, fumed, cryst.-free 1, 3, 4<br />

60676-86-0 Silica, vitreous 1, 4, 8<br />

55465-40-2 Silicic acid, aluminum potassium sodium salt 4<br />

68037-74-1<br />

67762-90-7<br />

Siloxanes and silicones, di-Me, polymers with Me<br />

silsesquioxanes<br />

Siloxanes and Silicones, di-Me, reaction products with<br />

silica<br />

63148-52-7 Siloxanes and silicones, dimethyl, 4<br />

1<br />

4<br />

1, 4<br />

Table continued on next page<br />

4<br />

4<br />

222


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

5324-84-5 Sodium 1-octanesulfonate 3<br />

2492-26-4 Sodium 2-mercaptobenzothiolate 2<br />

127-09-3 Sodium acetate 1, 3, 4<br />

532-32-1 Sodium benzoate 3<br />

144-55-8 Sodium bicarbonate 1, 2, 3, 4, 7<br />

7631-90-5 Sodium bisulfite 1, 3, 4<br />

1333-73-9 Sodium borate 1, 4, 6, 7<br />

7789-38-0 Sodium bromate 1, 2, 4<br />

7647-15-6 Sodium bromide 1, 2, 3, 4, 7<br />

1004542-84-0 Sodium bromosulfamate 8<br />

68610-44-6 Sodium caprylamphopropionate 4<br />

497-19-8 Sodium carbonate 1, 2, 3, 4, 8<br />

7775-09-9 Sodium chlorate 1, 4<br />

7647-14-5 Sodium chloride <br />

7758-19-2 Sodium chlorite <br />

3926-62-3 Sodium chloroacetate 3<br />

68608-68-4 Sodium cocaminopropionate 1<br />

142-87-0 Sodium decyl sulfate 1<br />

527-07-1 Sodium D-gluconate 4<br />

126-96-5 Sodium diacetate 1, 4<br />

2893-78-9 Sodium dichloroisocyanurate 2<br />

151-21-3 Sodium dodecyl sulfate 8<br />

1, 2, 3, 4,<br />

5, 8<br />

1, 2, 3, 4,<br />

5, 8<br />

6381-77-7 Sodium erythorbate (1:1) 1, 3, 4, 8<br />

126-92-1 Sodium ethasulfate 1<br />

141-53-7 Sodium formate 2, 8<br />

7681-38-1 Sodium hydrogen sulfate 4<br />

1310-73-2 Sodium hydroxide <br />

1, 2, 3, 4,<br />

7, 8<br />

7681-52-9 Sodium hypochlorite 1, 2, 3, 4, 8<br />

7681-82-5 Sodium iodide 4<br />

8061-51-6 Sodium ligninsulfonate 2<br />

18016-19-8 Sodium maleate (1:x) 8<br />

7681-57-4 Sodium metabisulfite 1<br />

Table continued on next page<br />

223


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

7775-19-1 Sodium metaborate 3, 4 <br />

16800-11-6 Sodium metaborate dihydrate 1, 4 <br />

10555-76-7 Sodium metaborate tetrahydrate 1, 4, 8 <br />

6834-92-0 Sodium metasilicate 1, 2, 4 <br />

7631-99-4 Sodium nitrate 2<br />

7632-00-0 Sodium nitrite 1, 2, 4 <br />

137-20-2 Sodium N-methyl-N-oleoyltaurate 4<br />

142-31-4 Sodium octyl sulfate 1<br />

1313-59-3 Sodium oxide 1<br />

11138-47-9 Sodium perborate 4<br />

10486-00-7 Sodium perborate tetrahydrate 1, 4, 5, 8 <br />

7632-04-4 Sodium peroxoborate 1<br />

7775-27-1 Sodium persulfate <br />

1, 2, 3, 4, <br />

7, 8 <br />

7632-05-5 Sodium phosphate 1, 4 <br />

9084-06-4 Sodium polynaphthalenesulfonate 2<br />

7758-16-9 Sodium pyrophosphate 1, 2, 4 <br />

54-21-7 Sodium salicylate 1, 4 <br />

533-96-0 Sodium sesquicarbonate 1, 2 <br />

1344-09-8 Sodium silicate 1, 2, 4 <br />

9063-38-1 Sodium starch glycolate 2<br />

7757-82-6 Sodium sulfate 1, 2, 3, 4 <br />

7757-83-7 Sodium sulfite 2, 4, 8 <br />

540-72-7 Sodium thiocyanate 1, 4 <br />

7772-98-7 Sodium thiosulfate 1, 2, 3, 4 <br />

10102-17-7 Sodium thiosulfate, pentahydrate 1, 4 <br />

650-51-1 Sodium trichloroacetate 1, 4 <br />

1300-72-7 Sodium xylenesulfonate 1, 3, 4 <br />

10377-98-7 Sodium zirconium lactate 1, 4 <br />

64742-88-7 Solvent naphtha (petroleum), medium aliph. 1, 2, 4 <br />

64742-96-7 Solvent naphtha, petroleum, heavy aliph. 2, 8 <br />

64742-94-5 Solvent naphtha, petroleum, heavy arom. 1, 2, 4, 5, 8 <br />

64742-95-6 Solvent naphtha, petroleum, light arom. 1, 2, 4 <br />

8007-43-0 Sorbitan, (9Z)-9-octadecenoate (2:3) 4<br />

Table continued on next page<br />

224


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

1338-43-8 Sorbitan, mono-(9Z)-9-octadecenoate 1, 2, 3, 4 <br />

9005-65-6<br />

9005-67-8<br />

Sorbitan, mono-(9Z)-9-octadecenoate, poly(oxy-1,2­<br />

ethanediyl) derivis.<br />

Sorbitan, monooctadecenoate, poly(oxy-1,2-ethanediyl) <br />

derivis.<br />

26266-58-0 Sorbitan, tri-(9Z)-9-octadecenoate 8<br />

<br />

<br />

3, 4 <br />

3, 4 <br />

10025-69-1 Stannous chloride dihydrate 1, 4 <br />

9005-25-8 Starch 1, 2, 4 <br />

68131-87-3<br />

Steam cracked distillate, cyclodiene dimer,<br />

dicyclopentadiene polymer<br />

8052-41-3 Stoddard solvent 1, 3, 4 <br />

10476-85-4 Strontium chloride 4<br />

100-42-5 Styrene 2<br />

57-50-1 Sucrose 1, 2, 3, 4 <br />

5329-14-6 Sulfamic acid 1, 4 <br />

14808-79-8 Sulfate 1, 4 <br />

68201-64-9 Sulfomethylated quebracho 2<br />

68608-21-9 Sulfonic acids, C10-16-alkane, sodium salts 6<br />

68439-57-6<br />

Sulfonic acids, C14-16-alkane hydroxy and C14-16­<br />

alkene, sodium salts<br />

61789-85-3 Sulfonic acids, petroleum 1<br />

68608-26-4 Sulfonic acids, petroleum, sodium salts 3<br />

1<br />

1, 3, 4 <br />

7446-09-5 Sulfur dioxide 2, 4, 8 <br />

7664-93-9 Sulfuric acid 1, 2, 4, 7 <br />

68955-19-1 Sulfuric acid, mono-C12-18-alkyl esters, sodium salts 4<br />

68187-17-7 Sulfuric acid, mono-C6-10-alkyl esters, ammonium salts 1, 4, 8 <br />

14807-96-6 Talc 1, 3, 4, 6, 7 <br />

8002-26-4 Tall oil 4, 8 <br />

61791-36-4 Tall oil imidazoline 4<br />

68092-28-4 Tall oil, compound with diethanolamine 1<br />

65071-95-6 Tall oil, ethoxylated 4, 8 <br />

8016-81-7 Tall-oil pitch 4<br />

61790-60-1 Tallow alkyl amines acetate 8<br />

72480-70-7<br />

Tar bases, quinoline derivatives, benzyl chloridequaternized<br />

1, 3, 4 <br />

68647-72-3 Terpenes and Terpenoids, sweet orange-oil 1, 3, 4, 8 <br />

Table continued on next page<br />

225


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

8000-41-7 Terpineol 1, 3 <br />

75-91-2 tert-Butyl hydroperoxide 1, 4 <br />

614-45-9 tert-Butyl perbenzoate 1<br />

12068-35-8 Tetra-calcium-alumino-ferrite 1, 2, 4 <br />

629-59-4 Tetradecane 8<br />

139-08-2 Tetradecyldimethylbenzylammonium chloride 1, 4, 8 <br />

112-60-7 Tetraethylene glycol 1, 4 <br />

112-57-2 Tetraethylenepentamine 1, 4 <br />

55566-30-8 Tetrakis(hydroxymethyl)phosphonium sulfate 1, 2, 3, 4, 7 <br />

681-84-5 Tetramethyl orthosilicate 1<br />

75-57-0 Tetramethylammonium chloride <br />

1, 2, 3, 4, <br />

7, 8 <br />

1762-95-4 Thiocyanic acid, ammonium salt 2, 3, 4 <br />

68-11-1 Thioglycolic acid 1, 2, 3, 4 <br />

62-56-6 Thiourea 1, 2, 3, 4, 6 <br />

68527-49-1<br />

Thiourea, polymer with formaldehyde and 1­<br />

phenylethanone<br />

68917-35-1 Thuja plicata donn ex. D. don leaf oil 3<br />

7772-99-8 Tin(II) chloride 1<br />

<br />

1, 4, 8 <br />

13463-67-7 Titanium dioxide 1, 2, 4 <br />

36673-16-2<br />

Titanium(4+) 2-[bis(2-hydroxyethyl)amino]ethanolate<br />

propan-2-olate (1:2:2)<br />

74665-17-1 Titanium, iso-Pr alc. triethanolamine complexes 1, 4 <br />

108-88-3 Toluene 1, 3, 4 <br />

126-73-8 Tributyl phosphate 1, 2, 4 <br />

81741-28-8 Tributyltetradecylphosphonium chloride 1, 3, 4 <br />

7758-87-4 Tricalcium phosphate 1, 4 <br />

12168-85-3 Tricalcium silicate 1, 2, 4 <br />

87-90-1 Trichloroisocyanuric acid 2<br />

629-50-5 Tridecane 8<br />

102-71-6 Triethanolamine 1, 2, 4 <br />

68299-02-5 Triethanolamine hydroxyacetate 3<br />

68131-71-5 Triethanolamine polyphosphate ester 1, 4, 8 <br />

77-93-0 Triethyl citrate 1, 4 <br />

78-40-0 Triethyl phosphate 1, 4 <br />

<br />

1<br />

Table continued on next page<br />

226


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

112-27-6 Triethylene glycol 1, 2, 3 <br />

112-24-3 Triethylenetetramine 4<br />

122-20-3 Triisopropanolamine 1, 4 <br />

14002-32-5 Trimethanolamine 3<br />

121-43-7 Trimethyl borate 8<br />

25551-13-7 Trimethylbenzene 1, 2, 4 <br />

7758-29-4 Triphosphoric acid, pentasodium salt 1, 4 <br />

1317-95-9 Tripoli 4<br />

6100-05-6 Tripotassium citrate monohydrate 4<br />

25498-49-1 Tripropylene glycol monomethyl ether 2<br />

68-04-2 Trisodium citrate 3<br />

6132-04-3 Trisodium citrate dihydrate 1, 4 <br />

150-38-9 Trisodium ethylenediaminetetraacetate 1, 3 <br />

19019-43-3 Trisodium ethylenediaminetriacetate 1, 4, 8 <br />

7601-54-9 Trisodium phosphate 1, 2, 4 <br />

10101-89-0 Trisodium phosphate dodecahydrate 1<br />

77-86-1 Tromethamine 3, 4 <br />

73049-73-7 Tryptone 8<br />

1319-33-1 Ulexite 1, 2, 3, 8 <br />

1120-21-4 Undecane 3, 8 <br />

57-13-6 Urea 1, 2, 4, 8 <br />

1318-00-9 Vermiculite 2<br />

24937-78-8 Vinyl acetate ethylene copolymer 1, 4 <br />

25038-72-6 Vinylidene chloride/methylacrylate copolymer 4<br />

7732-18-5 Water 2, 4, 8 <br />

8042-47-5 White mineral oil, petroleum 1, 2, 4 <br />

1330-20-7 Xylenes 1, 2, 4 <br />

8013-01-2 Yeast extract 8<br />

7440-66-6 Zinc 2<br />

3486-35-9 Zinc carbonate 2<br />

7646-85-7 Zinc chloride 1, 2 <br />

1314-13-2 Zinc oxide 1, 4 <br />

13746-89-9 Zirconium nitrate 2, 6 <br />

62010-10-0 Zirconium oxide sulfate 1, 4 <br />

Table continued on next page<br />

227


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

IUPAC<br />

CASRN Chemical Name Name and Reference<br />

Structure<br />

7699-43-6 Zirconium oxychloride 1, 2, 4<br />

21959-01-3 Zirconium(IV) chloride tetrahydrofuran complex 5<br />

14644-61-2 Zirconium(IV) sulfate 2, 6<br />

197980-53-3<br />

Zirconium, 1,1'-((2-((2-hydroxyethyl)(2­<br />

hydroxypropyl)amino)ethyl)imino)bis(2-propanol)<br />

complexes<br />

4<br />

68909-34-2 Zirconium, acetate lactate oxo ammonium complexes 4, 8<br />

174206-15-6 Zirconium, chloro hydroxy lactate oxo sodium complexes 4<br />

113184-20-6 Zirconium, hydroxylactate sodium complexes 1, 4<br />

101033-44-7<br />

Zirconium,tetrakis[2-[bis(2-hydroxyethyl)amino­<br />

kN]ethanolato-kO]­<br />

1, 2, 4, 5<br />

228


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table A-2 lists generic names of chemicals reported to be used in hydraulic fracturing fluids<br />

between 2005 and 2009. Generic chemical names provide limited information on the chemical, but<br />

are not specific enough to determine chemical structures. In some cases, the generic chemical name<br />

masks a specific chemical name and CASRN provided to the EPA and claimed as CBI by one or more<br />

of the nine hydraulic fracturing service companies.<br />

Table A-2. List of generic names of chemicals reportedly used in hydraulic fracturing fluids. In some cases, the<br />

generic chemical name masks a specific chemical name and CASRN provided to the EPA and claimed as CBI by<br />

one or more of the nine hydraulic fracturing service companies.<br />

Generic Chemical Name<br />

Reference<br />

2-Substituted aromatic amine salt 1, 4<br />

Acetylenic alcohol 1<br />

Acrylamide acrylate copolymer 4<br />

Acrylamide copolymer 1, 4<br />

Acrylamide modified polymer 4<br />

Acrylamide-sodium acrylate copolymer 4<br />

Acrylate copolymer 1<br />

Acrylic copolymer 1<br />

Acrylic polymer 1, 4<br />

Acrylic resin 4<br />

Acyclic hydrocarbon blend 1, 4<br />

Acylbenzylpyridinium choride 8<br />

Alcohol alkoxylate 1, 4<br />

Alcohol and fatty acid blend 2<br />

Alcohol ethoxylates 4<br />

Alcohols 1, 4<br />

Alcohols, C9-C22 1, 4<br />

Aldehydes 1, 4, 5<br />

Alfa-alumina 1, 4<br />

Aliphatic acids 1, 2, 3, 4<br />

Aliphatic alcohol 2<br />

Aliphatic alcohol glycol ether 3, 4<br />

Aliphatic alcohols, ethoxylated 2<br />

Aliphatic amine derivative 1<br />

Aliphatic carboxylic acid 4<br />

Alkaline bromide salts 1, 4<br />

Alkaline metal oxide 4<br />

Alkanes/alkenes 4<br />

Alkanolamine derivative 2<br />

Alkanolamine/aldehyde condensate 1, 2, 4<br />

Table continued on next page<br />

229


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

Generic Chemical Name<br />

Reference<br />

Alkenes 1, 4<br />

Alklaryl sulfonic acid 1, 4<br />

Alkoxylated alcohols 1<br />

Alkoxylated amines 1, 4<br />

Alkyaryl sulfonate 1, 2, 3, 4<br />

Alkyl alkoxylate 1, 4<br />

Alkyl amide 4<br />

Alkyl amine 1, 4<br />

Alkyl amine blend in a metal salt solution 1, 4<br />

Alkyl aryl amine sulfonate 4<br />

Alkyl aryl polyethoxy ethanol 3, 4<br />

Alkyl dimethyl benzyl ammonium chloride 4<br />

Alkyl esters 1, 4<br />

Alkyl ether phosphate 4<br />

Alkyl hexanol 1, 4<br />

Alkyl ortho phosphate ester 1, 4<br />

Alkyl phosphate ester 1, 4<br />

Alkyl phosphonate 4<br />

Alkyl pyridines 2<br />

Alkyl quaternary ammonium chlorides 1, 4<br />

Alkyl quaternary ammonium salt 4<br />

Alkylamine alkylaryl sulfonate 4<br />

Alkylamine salts 2<br />

Alkylaryl sulfonate 1, 4<br />

Alkylated quaternary chloride 1, 2, 4<br />

Alkylated sodium naphthalenesulphonate 2<br />

Alkylbenzenesulfonate 2<br />

Alkylbenzenesulfonic acid 1, 4, 5<br />

Alkylethoammonium sulfates 1<br />

Alkylphenol ethoxylates 1, 4<br />

Alkylpyridinium quaternary 4<br />

Alphatic alcohol polyglycol ether 2<br />

Aluminum oxide 1, 4<br />

Amide 4<br />

Amidoamine 1, 4<br />

Amine 1, 4<br />

Table continued on next page<br />

230


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

Generic Chemical Name<br />

Reference<br />

Amine compound 4<br />

Amine oxides 1, 4<br />

Amine phosphonate 1, 4<br />

Amine salt 1<br />

Amino compounds 1, 4<br />

Amino methylene phosphonic acid salt 1, 4<br />

Ammonium alcohol ether sulfate 1, 4<br />

Ammonium salt 1, 4<br />

Ammonium salt of ethoxylated alcohol sulfate 1, 4<br />

Amorphous silica 4<br />

Amphoteric surfactant 2<br />

Anionic acrylic polymer 2<br />

Anionic copolymer 1, 4<br />

Anionic polyacrylamide 1, 2, 4<br />

Anionic polyacrylamide copolymer 1, 4, 6<br />

Anionic polymer 1, 3, 4<br />

Anionic surfactants 2, 4, 6<br />

Antifoulant 1, 4<br />

Antimonate salt 1, 4<br />

Aqueous emulsion of diethylpolysiloxane 2<br />

Aromatic alcohol glycol ether 1<br />

Aromatic aldehyde 1, 4<br />

Aromatic hydrocarbons 3, 4<br />

Aromatic ketones 1, 2, 3, 4<br />

Aromatic polyglycol ether 1<br />

Arsenic compounds 4<br />

Ashes, residues 4<br />

Bentone clay 4<br />

Biocide 4<br />

Biocide component 1, 4<br />

Bis-quaternary methacrylamide monomer 4<br />

Blast furnace slag 4<br />

Borate salts 1, 2, 4<br />

Cadmium compounds 4<br />

Carbohydrates 1, 2, 4<br />

Carboxylmethyl hydroxypropyl guar 4<br />

Table continued on next page<br />

231


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

Generic Chemical Name<br />

Reference<br />

Cationic polyacrylamide 4<br />

Cationic polymer 2, 4<br />

Cedar fiber, processed 2<br />

Cellulase enzyme 1<br />

Cellulose derivative 1, 2, 4<br />

Cellulose ether 2<br />

Cellulosic polymer 2<br />

Ceramic 4<br />

Chlorous ion solution 1<br />

Chromates 1, 4<br />

Chrome-free lignosulfonate compound 2<br />

Citrus rutaceae extract 4<br />

Common white 4<br />

Complex alkylaryl polyo-ester 1<br />

Complex aluminum salt 1, 4<br />

Complex carbohydrate 2<br />

Complex organometallic salt 1<br />

Complex polyamine salt 7<br />

Complex substituted keto-amine 1<br />

Complex substituted keto-amine hydrochloride 1<br />

Copper compounds 6<br />

Coric oxide 4<br />

Cotton dust (raw) 2<br />

Cottonseed hulls 2<br />

Cured acrylic resin 1, 4<br />

Cured resin 1, 4, 5<br />

Cured urethane resin 1, 4<br />

Cyclic alkanes 1, 4<br />

Defoamer 4<br />

Dibasic ester 4<br />

Dicarboxylic acid 1, 4<br />

Diesel 1, 4, 6<br />

Dimethyl silicone 1, 4<br />

Dispersing agent 1<br />

Emulsifier 4<br />

Enzyme 4<br />

Table continued on next page<br />

232


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

Generic Chemical Name<br />

Reference<br />

Epoxy 4<br />

Epoxy resin 1, 4<br />

Essential oils 1, 4<br />

Ester Salt 2, 4<br />

Esters 2, 4<br />

Ether compound 4<br />

Ether salt 4<br />

Ethoxylated alcohol blend 4<br />

Ethoxylated alcohol/ester mixture 4<br />

Ethoxylated alcohols 1, 2, 4, 5, 7<br />

Ethoxylated alkyl amines 1, 4<br />

Ethoxylated amine blend 4<br />

Ethoxylated amines 1, 4<br />

Ethoxylated fatty acid 4<br />

Ethoxylated fatty acid ester 1, 4<br />

Ethoxylated nonionic surfactant 1, 4<br />

Ethoxylated nonylphenol 1, 2, 4<br />

Ethoxylated sorbitol esters 1, 4<br />

Ethylene oxide-nonylphenol polymer 4<br />

Fatty acid amine salt mixture 4<br />

Fatty acid ester 1, 2, 4<br />

Fatty acid tall oil 1, 4<br />

Fatty acids 1<br />

Fatty acid, ethoxylate 4<br />

Fatty alcohol alkoxylate 1, 4<br />

Fatty alkyl amine salt 1, 4<br />

Fatty amine carboxylates 1, 4<br />

Fatty imidazoline 4<br />

Fluoroaliphatic polymeric esters 1, 4<br />

Formaldehyde polymer 1<br />

Glass fiber 1, 4<br />

Glyceride esters 2<br />

Glycol 4<br />

Glycol blend 2<br />

Glycol ethers 1, 4, 7<br />

Ground cedar 2<br />

Table continued on next page<br />

233


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

Generic Chemical Name<br />

Reference<br />

Ground paper 2<br />

Guar derivative 1, 4<br />

Guar gum 4<br />

Haloalkyl heteropolycycle salt 1, 4<br />

Hexanes 1<br />

High molecular weight polymer 2<br />

High pH conventional enzymes 2<br />

Hydrocarbons 1<br />

Hydrogen solvent 4<br />

Hydrotreated and hydrocracked base oil 1, 4<br />

Hydrotreated distillate, light C9-16 4<br />

Hydrotreated heavy naphthalene 5<br />

Hydrotreated light distillate 2, 4<br />

Hydrotreated light petroleum distillate 4<br />

Hydroxyalkyl imino carboxylic sodium salt 2<br />

Hydroxycellulose 6<br />

Hydroxyethyl cellulose 1, 2, 4<br />

Imidazolium compound 4<br />

Inner salt of alkyl amines 1, 4<br />

Inorganic borate 1, 4<br />

Inorganic chemical 4<br />

Inorganic particulate 1, 4<br />

Inorganic salt 2, 4<br />

Iso-alkanes/n-alkanes 1, 4<br />

Isomeric aromatic ammonium salt 1, 4<br />

Latex 2, 4<br />

Lead compounds 4<br />

Low toxicity base oils 1, 4<br />

Lubra-Beads course 4<br />

Maghemite 1, 4<br />

Magnetite 1, 4<br />

Metal salt 1<br />

Metal salt solution 1<br />

Mineral 1, 4<br />

Mineral fiber 2<br />

Mineral filler 1<br />

Table continued on next page<br />

234


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

Generic Chemical Name<br />

Reference<br />

Mineral oil 4<br />

Mixed titanium ortho ester complexes 1, 4<br />

Modified acrylamide copolymer 2, 4<br />

Modified acrylate polymer 4<br />

Modified alkane 1, 4<br />

Modified bentonite 4<br />

Modified cycloaliphatic amine adduct 1, 4<br />

Modified lignosulfonate 2, 4<br />

Naphthalene derivatives 1, 4<br />

Neutralized alkylated napthalene sulfonate 4<br />

Nickel chelate catalyst 4<br />

Nonionic surfactant 1<br />

N-tallowalkyltrimethylenediamines 4<br />

Nuisance particulates 1, 2, 4<br />

Nylon 4<br />

Olefinic sulfonate 1, 4<br />

Olefins 1, 4<br />

Organic acid salt 1, 4<br />

Organic acids 1, 4<br />

Organic alkyl amines 4<br />

Organic chloride 4<br />

Organic modified bentonite clay 4<br />

Organic phosphonate 1, 4<br />

Organic phosphonate salts 1, 4<br />

Organic phosphonic acid salts 1, 4<br />

Organic polymer 4<br />

Organic polyol 4<br />

Organic salt 1, 4<br />

Organic sulfur compound 1, 4<br />

Organic surfactants 1<br />

Organic titanate 1, 4<br />

Organo amino silane 4<br />

Organo phosphonic acid 4<br />

Organo phosphonic acid salt 4<br />

Organometallic ammonium complex 1<br />

Organophilic clay 4<br />

Table continued on next page<br />

235


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

Generic Chemical Name<br />

Reference<br />

Oxidized tall oil 2<br />

Oxoaliphatic acid 2<br />

Oxyalkylated alcohol 1, 4<br />

Oxyalkylated alkyl alcohol 2, 4<br />

Oxyalkylated alkylphenol 1, 2, 3, 4<br />

Oxyalkylated fatty acid 1, 4<br />

Oxyalkylated fatty alcohol salt 2<br />

Oxyalkylated phenol 1, 4<br />

Oxyalkylated phenolic resin 4<br />

Oxyalkylated polyamine 1<br />

Oxyalkylated tallow diamine 2<br />

Oxyethylated alcohol 2<br />

Oxylated alcohol 1, 4<br />

P/F resin 4<br />

Paraffinic naphthenic solvent 1<br />

Paraffinic solvent 1, 4<br />

Paraffin inhibitor 4<br />

Paraffins 1<br />

Pecan shell 2<br />

Petroleum distallate blend 2, 3, 4<br />

Petroleum gas oils 1<br />

Petroleum hydrocarbons 4<br />

Petroleum solvent 2<br />

Phosphate ester 1, 4<br />

Phosphonate 2<br />

Phosphonic acid 1, 4<br />

Phosphoric acid, mixed polyoxyalkylene aryl and alkyl esters 4<br />

Plasticizer 1, 2<br />

Polyacrylamide copolymer 4<br />

Polyacrylamides 1<br />

Polyacrylate 1, 4<br />

Polyactide resin 4<br />

Polyalkylene esters 4<br />

Polyaminated fatty acid 2<br />

Polyaminated fatty acid surfactants 2<br />

Polyamine 1, 4<br />

Table continued on next page<br />

236


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

Generic Chemical Name<br />

Reference<br />

Polyamine polymer 4<br />

Polyanionic cellulose 1<br />

Polyaromatic hydrocarbons 6<br />

Polycyclic organic matter 6<br />

Polyelectrolyte 4<br />

Polyether polyol 2<br />

Polyethoxylated alkanol 2, 3, 4<br />

Polyethylene copolymer 4<br />

Polyethylene glycols 4<br />

Polyethylene wax 4<br />

Polyglycerols 2<br />

Polyglycol 2<br />

Polyglycol ether 6<br />

Polylactide resin 4<br />

Polymer 2, 4<br />

Polymeric hydrocarbons 3, 4<br />

Polymerized alcohol 4<br />

Polymethacrylate polymer 4<br />

Polyol phosphate ester 2<br />

Polyoxyalkylene phosphate 2<br />

Polyoxyalkylene sulfate 2<br />

Polyoxyalkylenes 1, 4, 7<br />

Polyphenylene ether 4<br />

Polyphosphate 4<br />

Polypropylene glycols 2<br />

Polyquaternary amine 4<br />

Polysaccaride polymers in suspension 2<br />

Polysaccharide 4<br />

Polysaccharide blend 4<br />

Polyvinylalcohol/polyvinylactetate copolymer 4<br />

Potassium chloride substitute 4<br />

Quarternized heterocyclic amines 4<br />

Quaternary amine 2, 4<br />

Quaternary amine salt 4<br />

Quaternary ammonium chloride 4<br />

Quaternary ammonium compound 1, 2, 4<br />

Table continued on next page<br />

237


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

Generic Chemical Name<br />

Reference<br />

Quaternary ammonium salts 1, 2, 4<br />

Quaternary compound 1, 4<br />

Quaternary salt 1, 4<br />

Quaternized alkyl nitrogenated compd 4<br />

Red dye 4<br />

Refined mineral oil 2<br />

Resin 4<br />

Salt of amine-carbonyl condensate 3, 4<br />

Salt of fatty acid/polyamine reaction product 3, 4<br />

Salt of phosphate ester 1<br />

Salt of phosphono-methylated diamine 1, 4<br />

Salts 4<br />

Salts of oxyalkylated fatty amines 4<br />

Sand 4<br />

Sand, AZ silica 4<br />

Sand, brown 4<br />

Sand, sacked 4<br />

Sand, white 4<br />

Secondary alcohol 1, 4<br />

Silica sand, 100 mesh, sacked 4<br />

Silicone emulsion 1<br />

Silicone ester 4<br />

Sodium acid pyrophosphate 4<br />

Sodium calcium magnesium polyphosphate 4<br />

Sodium phosphate 4<br />

Sodium salt of aliphatic amine acid 2<br />

Sodium xylene sulfonate 4<br />

Softwood dust 2<br />

Starch blends 6<br />

Substituted alcohol 1, 2, 4<br />

Substituted alkene 1<br />

Substituted alklyamine 1, 4<br />

Substituted alkyne 4<br />

Sulfate 4<br />

Sulfomethylated tannin 2, 5<br />

Sulfonate 4<br />

Table continued on next page<br />

238


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

Generic Chemical Name<br />

Reference<br />

Sulfonate acids 1<br />

Sulfonate surfactants 1<br />

Sulfonated asphalt 2<br />

Sulfonic acid salts 1, 4<br />

Sulfur compound 1, 4<br />

Sulphonic amphoterics 4<br />

Sulphonic amphoterics blend 4<br />

Surfactant blend 3, 4<br />

Surfactants 1, 2, 4<br />

Synthetic copolymer 2<br />

Synthetic polymer 4<br />

Tallow soap 4<br />

Telomer 4<br />

Terpenes 1, 4<br />

Titanium complex 4<br />

Triethanolamine zirconium chelate 1 4<br />

Triterpanes 4<br />

Vanadium compounds 4<br />

Wall material 1<br />

Walnut hulls 1, 2, 4<br />

Zirconium complex 2, 4<br />

Zirconium salt 4<br />

239


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table A-3 contains a list of chemicals with CASRNs that have been detected in flowback and<br />

produced water (collectively referred to as “hydraulic fracturing wastewater”). The table identifies<br />

chemicals that are also reported to be used in hydraulic fracturing fluids (Table A-1).<br />

Table A-3. List of CASRNs and names of chemicals detected in hydraulic fracturing wastewater. Chemicals also<br />

reportedly used in hydraulic fracturing fluids are marked with an “.”<br />

CASRN Chemical Name<br />

Also Listed in<br />

Table A - 1<br />

Reference<br />

87-61-6 1,2,3-Trichlorobenzene 3, 9<br />

120-82-1 1,2,4-Trichlorobenzene 9<br />

95-63-6 1,2,4-Trimethylbenzene 3, 9, 10<br />

57-55-6 1,2-Propanediol 3, 9<br />

108-67-8 1,3,5-Trimethylbenzene 3, 9, 10<br />

123-91-1 1,4-Dioxane 9, 10<br />

105-67-9 2,4-Dimethylphenol 3, 9, 10<br />

87-65-0 2,6-Dichlorophenol 3, 9<br />

91-57-6 2-Methylnaphthalene 3, 9, 10<br />

95-48-7 2-Methylphenol 3, 9, 10<br />

79-31-2 2-Methylpropanoic acid 10<br />

109-06-8 2-Methylpyridine 3, 9<br />

503-74-2 3-Methylbutanoic acid 10<br />

108-39-4 3-Methylphenol 3, 9, 10<br />

106-44-5 4-Methylphenol 3, 9, 10<br />

57-97-6 7,12-Dimethylbenz(a)anthracene 3, 9<br />

64-19-7 Acetic acid 3, 9, 10<br />

67-64-1 Acetone 3, 9, 10<br />

98-86-2 Acetophenone 3, 9<br />

107-02-8 Acrolein 9<br />

107-13-1 Acrylonitrile 3, 9<br />

309-00-2 Aldrin 3, 9<br />

7429-90-5 Aluminum 3, 9, 10<br />

7664-41-7 Ammonia 3, 9, 10<br />

7440-36-0 Antimony 3, 9, 10<br />

12672-29-6 Aroclor 1248 3, 9<br />

7440-38-2 Arsenic 3, 9, 10<br />

7440-39-3 Barium 3, 9, 10<br />

71-43-2 Benzene 3, 9, 10<br />

50-32-8 Benzo(a)pyrene 3, 9<br />

205-99-2 Benzo(b)fluoranthene 3, 9<br />

191-24-2 Benzo(g,h,i)perylene 3, 9, 10<br />

207-08-9 Benzo(k)fluoranthene 3, 9<br />

100-51-6 Benzyl alcohol 3, 9, 10<br />

Table continued on next page<br />

240


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

CASRN<br />

7440-41-7<br />

Chemical Name<br />

Beryllium<br />

Also Listed in<br />

Table A -1 <br />

319-85-7 beta-1,2,3,4,5,6-Hexachlorocyclohexane 3, 9 <br />

111-44-4 Bis(2-chloroethyl) ether 3, 9 <br />

Reference<br />

3, 9, 10 <br />

7440-42-8 Boron 3, 9, 10 <br />

24959-67-9 Bromide (-1) 3, 9, 10 <br />

75-27-4 Bromodichloromethane 3<br />

75-25-2 <br />

107-92-6 <br />

Bromoform<br />

Butanoic acid<br />

3, 9, 10 <br />

9, 10 <br />

104-51-8<br />

Butylbenzene 9, 10 <br />

7440-43-9 Cadmium 3, 9, 10 <br />

10045-97-3 Caesium 137 3<br />

7440-70-2 Calcium 3, 9, 10 <br />

124-38-9 Carbon dioxide 3, 9, 10 <br />

75-15-0 Carbon disulfide 3, 9 <br />

16887-00-6 Chloride <br />

3, 9, 10 <br />

7782-50-5 Chlorine 3, 10 <br />

124-48-1 Chlorodibromomethane 3<br />

67-66-3 Chloroform 3, 9, 10 <br />

74-87-3 Chloromethane 3, 10 <br />

7440-47-3 Chromium 3, 9, 10 <br />

16065-83-1 Chromium (III), insoluble salts 3<br />

18540-29-9 Chromium (VI) 3, 10 <br />

7440-48-4 Cobalt 3, 9, 10 <br />

7440-50-8 Copper <br />

98-82-8 Cumene 3, 9 <br />

3, 9, 10 <br />

57-12-5 Cyanide, free 3, 9, 10 <br />

319-86-8 delta-Hexachlorocyclohexane 9<br />

117-81-7 Di(2-ethylhexyl) phthalate <br />

53-70-3 Dibenz(a,h)anthracene 3, 9 <br />

3, 9, 10 <br />

84-74-2 Dibutyl phthalate 3, 9, 10 <br />

75-09-2 Dichloromethane 9, 10 <br />

60-57-1 Dieldrin 9<br />

84-66-2 Diethyl phthalate 9<br />

117-84-0 <br />

122-39-4 <br />

Dioctyl phthalate<br />

Diphenylamine<br />

959-98-8 Endosulfan I 3, 9 <br />

33213-65-9 Endosulfan II 3, 9 <br />

7421-93-4 Endrin aldehyde 3, 9 <br />

9, 10 <br />

3, 9 <br />

Table continued on next page<br />

241


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

CASRN<br />

100-41-4<br />

107-21-1 <br />

Chemical Name<br />

Ethylbenzene<br />

Also Listed in<br />

Table A -1 <br />

Ethylene glycol 3, 9 <br />

<br />

Reference<br />

3, 9, 10 <br />

206-44-0<br />

Fluoranthene 3, 9 <br />

86-73-7 Fluorene 3, 9, 10 <br />

16984-48-8 Fluoride 3, 9, 10 <br />

64-18-6 Formic acid 10 <br />

76-44-8 Heptachlor 3, 9 <br />

1024-57-3 Heptachlor epoxide 3, 9 <br />

111-14-8 <br />

Heptanoic acid<br />

142-62-1<br />

Hexanoic acid 10 <br />

193-39-5 Indeno(1,2,3-cd)pyrene 3, 9 <br />

7439-89-6 Iron <br />

67-63-0 <br />

Isopropanol 3, 9 <br />

7439-92-1 Lead <br />

58-89-9 Lindane 3, 9 <br />

10 <br />

3, 9, 10 <br />

3, 9, 10 <br />

7439-93-2 Lithium 3, 9, 10 <br />

7439-95-4 Magnesium 3, 9, 10 <br />

7439-96-5 Manganese 3, 9, 10 <br />

7439-97-6 <br />

67-56-1 <br />

Mercury<br />

Methanol 3, 9 <br />

74-83-9 Methyl bromide 3, 9 <br />

3, 9, 10 <br />

78-93-3 Methyl ethyl ketone 3, 9, 10 <br />

7439-98-7 Molybdenum 3, 9, 10 <br />

91-20-3 Naphthalene <br />

7440-02-0 <br />

Nickel<br />

86-30-6 N-Nitrosodiphenylamine 3, 9 <br />

72-55-9 p,p'-DDE 3, 9 <br />

3, 9, 10 <br />

3, 9, 10 <br />

99-87-6 p-Cymene 9, 10 <br />

109-52-4 <br />

Pentanoic acid<br />

85-01-8 Phenanthrene <br />

108-95-2 <br />

298-02-2<br />

Phorate 9<br />

10 <br />

3, 9, 10 <br />

Phenol 3, 9, 10 <br />

7723-14-0 Phosphorus 3, 9 <br />

7440-09-7 Potassium 3, 9, 10 <br />

79-09-4 Propionic acid 10 <br />

103-65-1 <br />

Propylbenzene<br />

129-00-0<br />

Pyrene 9, 10 <br />

110-86-1 Pyridine 3, 9, 10 <br />

9<br />

Table continued on next page<br />

242


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

CASRN<br />

13982-63-3<br />

Chemical Name<br />

Radium 226<br />

Also Listed in<br />

Table A -1<br />

7440-14-4 Radium 226,228 3<br />

Reference<br />

3, 10<br />

15262-20-1 Radium 228 3, 10<br />

94-59-7 Safrole 3, 9<br />

135-98-8 sec-Butylbenzene 9<br />

7782-49-2 Selenium 3, 9, 10<br />

7631-86-9 Silica 10<br />

7440-21-3 Silicon (elemental) 10<br />

7440-22-4 Silver 3, 9, 10<br />

7440-23-5 Sodium 3, 9, 10<br />

7440-24-6 Strontium 3, 9, 10<br />

14808-79-8 Sulfate <br />

14265-45-3 Sulfite 3<br />

127-18-4 Tetrachloroethylene 3, 9<br />

3, 9, 10<br />

7440-28-0 Thallium and Compounds 3, 9, 10<br />

7440-31-5 Tin 9, 10<br />

7440-32-6 Titanium 3, 9, 10<br />

108-88-3 Toluene <br />

3, 9, 10<br />

7440-62-2 Vanadium 3, 10<br />

1330-20-7 Xylenes <br />

7440-66-6 Zinc <br />

7440-67-7 Zirconium 3<br />

3, 9, 10<br />

3, 9, 10<br />

243


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table A-4 contains a list of chemicals and properties that are detected in flowback and produced<br />

water (collectively referred to as “hydraulic fracturing wastewater”).<br />

Table A-4. List of chemicals and properties detected in hydraulic fracturing wastewater.<br />

Chemical Name / Property Reference<br />

Alkalinity 3, 9, 10<br />

Alkalinity, carbonate (as CaCO 3 ) 3, 9, 10<br />

Alpha radiation 3<br />

Aluminum, dissolved 3, 9<br />

Barium strontium P.S. 3<br />

Barium, dissolved 3, 9<br />

Beta radiation 3<br />

Bicarbonates (HCO 3 ) 3, 10<br />

Biochemical oxygen demand 3, 9, 10<br />

Cadmium, dissolved 3, 9<br />

Calcium, dissolved 3, 9<br />

Chemical oxygen demand 3, 9, 10<br />

Chromium (VI), dissolved 3<br />

Chromium, dissolved 3, 9<br />

Cobalt, dissolved 3, 9<br />

Coliform 3<br />

Color 3<br />

Conductivity 3, 9, 10<br />

Hardness as CaCO 3 3, 9, 10<br />

Heterotrophic plate count 3<br />

Hexanoic acid 10<br />

Iron, dissolved 3, 9<br />

Lithium, dissolved 3, 9<br />

Magnesium, dissolved 3, 9<br />

Chemical Name / Property Reference<br />

Manganese, dissolved 3, 9<br />

Nickel, dissolved 3, 9<br />

Nitrate, as N 3, 9, 10<br />

Nitrogen, total as N 3<br />

Oil and grease 3, 9, 10<br />

Petroleum hydrocarbons 3<br />

pH 3, 9, 10<br />

Phenols 3<br />

Potassium, dissolved 3, 9<br />

Salt 3<br />

Scale inhibitor 3<br />

Selenium, dissolved 3, 9<br />

Silver, dissolved 3, 10<br />

Sodium, dissolved 3, 10<br />

Strontium, dissolved 3, 10<br />

Surfactants 3<br />

Total alkalinity 3, 9, 10<br />

Total dissolved solids 3, 9, 10<br />

Total Kjeldahl nitrogen 3, 9, 10<br />

Total organic carbon 3, 9, 10<br />

Total sulfide 9<br />

Total suspended solids 3, 9, 10<br />

Volatile acids 3, 9<br />

Zinc, dissolved 3, 9<br />

244


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

References<br />

1. US House of Representatives 2011. Chemicals Used in Hydraulic Fracturing. Available at<br />

http://democrats.energycommerce.house.gov/sites/default/files/documents/Hydraulic%2<br />

0Fracturing%20Report%204.18.11.pdf. Accessed November 27, 2012.<br />

2. Colborn, T., Kwiatkowski, C., Schultz, K. and Bachran, M. 2011. Natural Gas Operations from<br />

a Public Health Perspective. Human and Ecological Risk Assessment 17 (5): 1039-1056.<br />

3. New York State Department of Environmental Conservation. 2011. Supplemental Generic<br />

Environmental Impact Statement on the Oil, Gas and Solution Mining Regulatory Program<br />

(Revised Draft). Well Permit Issuance for Horizontal Drilling and High-Volume Hydraulic<br />

Fracturing to Develop the Marcellus Shale and Other Low-Permeability Gas Reservoirs.<br />

Available at http://www.dec.ny.gov/energy/75370.html. Accessed September 1, 2011.<br />

4. US Environmental Protection Agency. 2011. Data received from oil and gas exploration and<br />

production companies, including hydraulic fracturing service companies. Non-confidential<br />

business information source documents are located in Federal Docket ID: EPA-HQ-ORD­<br />

2010-0674. Available at http://www.regulations.gov. Accessed November 14, 2012.<br />

5. Material Safety Data Sheets. (a) Encana/Halliburton Energy Services, Inc.: Duncan,<br />

Oklahoma. Provided by Halliburton Energy Services during an onsite visit by the EPA on<br />

May 10, 2010; (b) Encana Oil and Gas (USA), Inc.: Denver, Colorado. Provided to US EPA<br />

Region 8.<br />

6. US Environmental Protection Agency, Office of Water. 2004. Evaluation of Impacts to<br />

Underground Sources of Drinking Water by Hydraulic Fracturing of Coalbed Methane<br />

Reservoirs. EPA 816-R-04-003. Available at http://water.epa.gov/type/groundwater/uic/<br />

class2/hydraulicfracturing/wells_coalbedmethanestudy.cfm. Accessed November 27, 2012.<br />

7. Pennsylvania Department of Environmental Protection. 2010. Chemicals Used by Hydraulic<br />

Fracturing Companies in Pennsylvania for Surface and Hydraulic Fracturing Activities.<br />

Available at http://files.dep.state.pa.us/OilGas/BOGM/BOGMPortalFiles/MarcellusShale/<br />

Frac%20list%206-30-2010.pdf. Accessed November 27, 2012.<br />

8. Ground Water Protection Council. 2012. FracFocus well records: January 1, 2011, through<br />

February 27, 2012. Available at http://www.fracfocus.org/.<br />

9. Hayes, T. 2009. Sampling and Analysis of Water Streams Associated with the Development<br />

of Marcellus Shale Gas. Gas Technology Institute for Marcellus Shale Coalition. Available at<br />

http://eidmarcellus.org/wp-content/uploads/2012/11/MSCommission-Report.pdf.<br />

Accessed November 30, 2012.<br />

10. US Environmental Protection Agency. 2011. Sampling Data for Flowback and Produced<br />

Water Provided to EPA by Nine Oil and Gas Well Operators (Non-Confidential Business<br />

Information). Available at http://www.regulations.gov/#!docketDetail;rpp=100;so=DESC;<br />

sb=docId;po=0;D=EPA-HQ-ORD-2010-0674. Accessed November 27, 2012.<br />

245


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Appendix B: Stakeholder Engagement 90<br />

B.1. Stakeholder Engagement Road Map for the EPA’s Study on the Potential<br />

Impacts of Hydraulic Fracturing on Drinking Water Resources<br />

On March 18, 2010, at the request of the U.S. Congress, the EPA announced plans to develop a<br />

comprehensive research study on the potential impact of hydraulic fracturing on drinking water<br />

resources. The EPA believes a transparent, research-driven approach with significant stakeholder<br />

involvement can address questions about hydraulic fracturing and strengthen our clean energy<br />

future. The road map below outlines the EPA’s plans to build upon its commitment to transparency<br />

and stakeholder engagement coordinated during the development of the Study Plan and will help<br />

inform the report of results.<br />

Goals of Strengthened Stakeholder Engagement<br />

• Increase technical engagement with the stakeholder community to ensure that the EPA<br />

has ongoing access to a broad range of expertise and data outside the agency.<br />

• Improve public understanding of the goals and design of the study.<br />

• Ensure that the EPA is current on changes in industry practices and technologies so the<br />

report of results reflects an up-to-date picture of hydraulic fracturing operations.<br />

• Obtain timely and constructive feedback on projects undertaken as part of the study.<br />

Increased Technical Engagement<br />

In November 2012, the EPA held five roundtables focused on each stage of the water cycle:<br />

• Water acquisition. This study takes steps to examine potential changes in the quantity of<br />

water available for drinking and potential changes in drinking water quality that result<br />

from acquisition for hydraulic fracturing. The EPA is aware that the use of recycling is<br />

rapidly growing and that this may affect the need to acquire water for hydraulic fracturing.<br />

• Chemical mixing. The study examines the potential release of chemicals used in hydraulic<br />

fracturing to surface and ground water through onsite spills and/or leaks and compiles<br />

information on hydraulic fracturing fluids and chemicals from publicly available data, data<br />

provided by nine hydraulic fracturing service companies and other sources.<br />

• Flowback. The study examines available data regarding release to surface or ground water<br />

through spills or leakage from onsite storage.<br />

• Water treatment and disposal. The study examines the potential for contaminants to reach<br />

drinking water due to surface water discharge, the effectiveness of current wastewater<br />

treatment, and the potential formation of disinfection byproducts in drinking water<br />

treatment facilities.<br />

90 The text and figure included in this appendix were taken from http://www.epa.gov/<strong>hf</strong>study/stakeholderroadmap.html.<br />

Please see this website for updated information as it becomes available.<br />

246


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

• Well injection. The study takes steps to examine the potential for release of hydraulic<br />

fracturing fluids to ground water due to inadequate well construction or operation,<br />

movement of hydraulic fracturing fluids from the target formation to drinking water<br />

aquifers through local man-made or natural features (e.g., other production or abandoned<br />

wells and existing faults or fractures).<br />

Based on feedback from these roundtables, the EPA will host in-depth technical workshops to<br />

address specific issues in greater detail. These technical workshops will begin in February 2013<br />

and continue as needed. Upon completion of the last technical workshop, the EPA will reconvene<br />

the original roundtables to review the work addressed in the technical workshop series.<br />

Improve Public Understanding<br />

To improve public understanding of the study, the EPA staff will increase the frequency of<br />

webinars. For instance, after the initial set of roundtables and each technical workshop, the EPA<br />

will host a webinar to report out to the public on these. The EPA will continue to provide regular<br />

electronic updates to its list of stakeholders.<br />

In addition to the webinars, the EPA staff will regularly update its hydraulic fracturing study<br />

website with up-to-date materials and identify opportunities for briefings and updates on the study<br />

to stakeholders (e.g., annual or regional meetings of industry trade associations, annual meetings of<br />

environmental/public health groups, academic conferences, annual or regional meetings of water<br />

utilities, and tribal meetings).<br />

The EPA has previously committed to the release in December 2012 of a progress report on the<br />

study. While the progress report will not make any final findings or conclusions, it will provide the<br />

public with an update on study activities and future work.<br />

Ensure the EPA is Current on Industry Practices<br />

To ensure that the EPA is up-to-date on evolving industry practices and technologies, the EPA will<br />

publish a Federal Register notice in late 2012 to create a docket where stakeholders can submit<br />

peer-reviewed data from ongoing or completed studies. This initial request will be followed up with<br />

requests in 2013 and 2014.<br />

Obtain Timely Feedback<br />

The EPA intends to receive timely feedback on the projects conducted as part of the study through<br />

the roundtables and technical workshops described above. In addition, the EPA's Science Advisory<br />

Board is forming a panel of independent experts who will provide advice and review under the<br />

auspices of the Science Advisory Board on the EPA's hydraulic fracturing research described in its<br />

2012 Progress Report. The EPA plans to use such advice for the development of a report of results,<br />

estimated to be released in late 2014, which will also be reviewed by the Science Advisory Board. In<br />

addition, this panel may also provide advice on other technical documents and issues related to<br />

hydraulic fracturing upon further request by the EPA. The panel will provide opportunities for<br />

public comment in connection with these activities.<br />

247


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

B.2 Stakeholder Road Map and Timeline<br />

Increase technical engagement with the stakeholder community to ensure that the EPA has<br />

ongoing access to a broad range of expertise and data outside the agency.<br />

Plan: The week of November 12, 2012, EPA held five roundtables focused on each stage of the water<br />

cycle, to be followed in Spring 2013 by a series of technical workshops on topics identified during<br />

the roundtables.<br />

Implementation:<br />

• Identify participants for meetings (September 2012):<br />

o<br />

o<br />

The EPA consulted with industry, non-governmental organizations, states, and<br />

tribes through a series of one-on-one meetings in September to present the plan<br />

for the roundtables and ask for potential invitees with technical expertise. The EPA<br />

then selected invitees with appropriate technical backgrounds.<br />

Roundtable participants numbered 15-20 in addition to the EPA staff.<br />

• Kick-off (October 2012)<br />

o<br />

The EPA hosted a kick-off (virtual) meeting with technical representatives<br />

representing a broad range of stakeholders to lay out the context, goals, and<br />

logistics for the roundtables.<br />

• Roundtables (November 14–16, 2012)<br />

o<br />

o<br />

Each meeting was professionally facilitated.<br />

All roundtables occurred in DC. These were half-day meetings.<br />

• Workshops (February 2013 through April 2013)<br />

• Second round of roundtables (Summer/Fall 2013)<br />

Obtain timely and constructive feedback on projects undertaken as part of the study and<br />

ensure that the EPA is current on changes in industry practices and technologies so the<br />

report of results reflects an up-to-date picture of hydraulic fracturing operations.<br />

Plan: Issue Federal Register notices in 2012, 2013, and 2014 requesting additional data and<br />

information to inform the study. 91 The notices will request peer-reviewed data and reports that can<br />

help answer the research questions, for example, the content of hydraulic fracturing flowback and<br />

produced water; the location of prior wastewater treatment pits, ponds, lagoons, and tanks; specific<br />

sources of water used for hydraulic fracturing; specific water quality requirements for use of water<br />

or reuse of waste water in hydraulic fracturing; partitioning of constituents into gas solid and liquid<br />

components (particularly the fate of metals, organics, and radionuclides).<br />

91 The first Federal Register notice was published in November 2012 and is available at http://www.gpo.gov/fdsys/<br />

pkg/FR-2012-11-09/pdf/2012-27452.pdf.<br />

248


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Implementation:<br />

• Technical workshops on specific technical topics suggested by roundtable participants<br />

(begin February 2013)<br />

• These sessions will flow from roundtable discussions. The EPA will convene experts to<br />

address specific issues of data collection, method or data interpretation (i.e., how to find<br />

more comprehensive/reliable spill data; how to get good data for the environmental<br />

justice analysis, etc.). The EPA will issue the first Federal Register notice in late 2012 to<br />

request peer-reviewed data and studies that can help answer the research questions.<br />

Additional Federal Register notices will request peer-reviewed information and will be<br />

published annually in 2013 and 2014.<br />

Improve public understanding of the goals and design of the study.<br />

Plan: In addition to the organized technical meetings, the EPA will seek opportunities (such as<br />

association or state organization meetings) to provide informal briefings and updates on the study<br />

to a diverse range of stakeholders, including states, non-governmental organizations, academia, and<br />

industry. The EPA will also increase the frequency of webinars, hosting them after each technical<br />

meeting to report out to the public on the discussion.<br />

Implementation: The EPA will host monthly webinars following the initial set of roundtables and<br />

each technical workshop to inform the public of topics discussed. The EPA will develop and publish<br />

a calendar of events where presentations on the study will be made.<br />

249


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Figure B-1. Timeline for technical roundtables and workshops. The goals of this enhanced engagement process are to improve public understanding of the study,<br />

ensure that the EPA is current on changes in industry practices and technologies so that the report of results reflects an up-to-date picture of hydraulic fracturing<br />

operations, and obtain timely and constructive feedback on ongoing research projects.<br />

250


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Appendix C: Summary of QAPPs<br />

This appendix provides a quick reference table for QAPPs associated with the research projects that<br />

comprise the EPA’s Study of the Potential Impacts of Drinking Water Resources. Current versions of<br />

the QAPPs are available at http://www.epa.gov/<strong>hf</strong>study/qapps.html.<br />

Table C-1. QAPPs associated with the research projects discussed in this progress report.<br />

Research Project<br />

Literature Review<br />

Spills Database Analysis<br />

Service Company Analysis<br />

Well File Review<br />

FracFocus Analysis<br />

Subsurface Migration<br />

Modeling<br />

Surface Water Modeling<br />

Water Availability Modeling<br />

Source Apportionment<br />

Studies<br />

Wastewater Treatability<br />

Studies<br />

Br-DBP Precursor Studies<br />

Analytical Method<br />

Development<br />

QAPP Title<br />

QAPP for Hydraulic Fracturing Data and Literature Evaluation for the<br />

EPA’s Study of the Potential Impacts of Hydraulic Fracturing on<br />

Drinking Water Resources<br />

QAPP for Hydraulic Fracturing Surface Spills Data Analysis<br />

Final QAPP for the Evaluation of Information on Hydraulic Fracturing<br />

QAPP for Analysis of Data Received from Nine Hydraulic Fracturing<br />

Service Companies<br />

QAPP for Hydraulic Fracturing<br />

National Hydraulic Fracturing Study Evaluation of Existing Production<br />

Well File Contents: QAPP<br />

Supplemental Programmatic QAPP for Work Assignment 4-58:<br />

National Hydraulic Fracturing Study Evaluation of Existing Production<br />

Well File Contents<br />

Supplemental Programmatic QAPP for Work Assignment 4-58:<br />

National Hydraulic Fracturing Study Evaluation of Existing Production<br />

Well File Contents<br />

QAPP for Analysis of Data Extracted from FracFocus<br />

Analysis of Environmental Hazards Related to Hydrofracturing<br />

QAPP for Surface Water Transport of Hydraulic Fracturing-Derived<br />

Waste Water<br />

Data Collection/Mining for Hydraulic Fracturing Case Studies<br />

Modeling the Impact of Hydraulic Fracturing on Water Resources<br />

Based on Water Acquisition Scenarios<br />

QAPP for Hydraulic Fracturing Waste Water Source Apportionment<br />

Study<br />

Fate, Transport, and Characterization of Contaminants in Hydraulic<br />

Fracturing Water in Wastewater Treatment Processes<br />

Formation of Disinfection By-Products from Hydraulic Fracturing Fluid<br />

Constituents: QAPP<br />

QAPP for the Chemical Characterization of Select Constituents<br />

Relevant to Hydraulic Fracturing<br />

QAPP for the Interlaboratory Verification and Validation of Diethylene<br />

Glycol, Triethylene Glycol, Tetraethylene Glycol, 2-Butoxyethanol and<br />

2-Methoxyethanol in Ground and Surface Waters by Liquid<br />

Chromatography/Tandem Mass Spectrometry<br />

Table continued on next page<br />

251


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Table continued from previous page<br />

Research Project<br />

QAPP Title<br />

QAPP: Health and Toxicity Theme, Hydraulic Fracturing Study<br />

Toxicity Assessment<br />

QAPP for Chemical Information Quality Review and Physicochemical<br />

Property Calculations for Hydraulic Fracturing Chemical Lists<br />

Las Animas and Huerfano<br />

Hydraulic Fracturing Retrospective Case Study, Raton Basin, CO<br />

Counties, Colorado<br />

Hydraulic Fracturing Retrospective Case Study, Bakken Shale,<br />

Dunn County, North Dakota<br />

Killdeer and Dunn County, ND<br />

Bradford County,<br />

Pennsylvania<br />

Washington County,<br />

Pennsylvania<br />

Wise County, Texas<br />

Hydraulic Fracturing Retrospective Case Study, Bradford-<br />

Susquehanna Counties, PA<br />

Hydraulic Fracturing Retrospective Case Study, Marcellus Shale,<br />

Washington County, PA<br />

Hydraulic Fracturing Retrospective Case Study, Wise and Denton<br />

Cos., TX<br />

252


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Appendix D: Divisions of Geologic Time<br />

Figure E-1 is reproduced from a USGS fact<br />

sheet, “ Divisions of<br />

Geologic Time:<br />

Major<br />

Chronostratigraphic and Geochronological<br />

Units.” A geologic timescale relates rock<br />

layers to time.<br />

Chronstratigraphic units refer to specific<br />

rock layers while geochronological units<br />

refer to specific time periods.<br />

Reference<br />

US Geological Survey. 2010. Divisions of<br />

Geologic Time: Major Chronostratigraphic<br />

and Geochronological Units. Fact Sheet<br />

2010-3059. US Geological Survey. Available<br />

at http://pubs.usgs.gov/fs/<br />

2010/3059/pdf/FS10-3059.pdf. Accessed<br />

November 30, 2012.<br />

Figure D-1. Divisions of geologic time approved by<br />

the USGS Geologic Names Committee (2010).The<br />

chart shows major chronostratigraphic and<br />

geochronologic units.<br />

253


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Glossary<br />

Acid mine drainage: Drainage of water from areas that have been mined for coal of other mineral<br />

ores. The water has a low pH because of its contact with sulfur-bearing material and is harmful to<br />

aquatic organisms. (2)<br />

Adsorption: Adhesion of molecules of gas, liquid, or dissolved solids to a surface. (2)<br />

Aeration: A process that promotes biological degradation of organic matter in water. The process<br />

may be passive (as when waste is exposed to air) or active (as when a mixing or bubbling device<br />

introduces the air). (2)<br />

Ambient water quality: Natural concentration of water quality constituents prior to mixing of<br />

either point or nonpoint source load of contaminants. Reference ambient concentration is used to<br />

indicate the concentration of a chemical that will not cause adverse impact to human health. (2)<br />

Analysis of existing data: The process of gathering and summarizing existing data from various<br />

sources to provide current information on hydraulic fracturing activities.<br />

Analyte: The element, ion, or compound that an analysis seeks to identify; the compound of<br />

interest. (2)<br />

Annulus: Either the space between the casing of a well and the wellbore or the space between the<br />

tubing and casing of a well. (2)<br />

API number: A unique identifying number for all oil and gas wells drilled in the United States. The<br />

system was developed by the American Petroleum Institute. (1)<br />

Aquifer: An underground geological formation, or group of formations, containing water. A source<br />

of ground water for wells and springs. (2)<br />

Baseline data: Initial information on a program or program components collected prior to receipt<br />

of services or participation activities. Often gathered through intake interviews and observations<br />

and used later for comparing measures that determine changes in a program. (2)<br />

Case study: An approach often used in research to better understand real-world situations or<br />

events using a systematic process for the collection and analysis of data.<br />

Prospective case study: A study of sites where hydraulic fracturing will occur after the<br />

research is initiated. These case studies allow sampling and characterization of the site<br />

prior to, and after, water extraction, drilling, hydraulic fracturing fluid injection, flowback,<br />

and gas production. The data collected during prospective case studies will allow the EPA to<br />

evaluate any changes in water quality over time.<br />

Retrospective case study: A study of sites where hydraulic fracturing has occurred nearby,<br />

with a focus on sites with reported instances of drinking water resource contamination.<br />

254


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

These studies will use existing data, sampling, and possibly modeling to determine whether<br />

reported impacts are due to hydraulic fracturing activities or other sources.<br />

Casing: Pipe cemented in the well to seal off formation fluids and to keep the hole from caving in.<br />

(1)<br />

Chemical Abstracts Service: Provides information on chemical properties and interactions. Every<br />

year, the Chemical Abstracts Service updates and writes new chemical abstracts on well over a<br />

million different chemicals, including each chemical’s composition, structure, characteristics, and<br />

different names. Each abstract is accompanied by a registration number, or CASRN. (2)<br />

Coalbed methane: Methane contained in coal seams. A coal seam is a layer or stratum of coal<br />

parallel to the rock stratification.<br />

Confidential business information (CBI): Information that contains trade secrets, commercial or<br />

financial information, or other information that has been claimed as confidential by the submitter.<br />

The EPA has special procedures for handling such information. (2)<br />

Contaminant: A substance that is either present in an environment where it does not belong or is<br />

present at levels that might cause harmful (adverse) health effects. (2)<br />

Conventional reservoir: A reservoir in which buoyant forces keep hydrocarbons in place below a<br />

sealing caprock. Reservoir and fluid characteristics of conventional reservoirs typically permit oil<br />

or natural gas to flow readily into wellbores. The term is used to make a distinction from shale and<br />

other unconventional reservoirs, in which gas might be distributed throughout the reservoir at the<br />

basin scale, and in which buoyant forces or the influence of a water column on the location of<br />

hydrocarbons within the reservoir are not significant. (5)<br />

Discharge: Any emission (other than natural seepage), intentional or unintentional. Includes, but is<br />

not limited to, spilling, leaking, pumping, pouring, emitting, emptying or dumping. (2)<br />

Disinfection byproduct (DBP): A compound formed by the reaction of a disinfectant such as<br />

chlorine with organic material in the water supply. (2)<br />

Drinking water resource: Any body of water, ground or surface, that could currently, or in the<br />

future, serve as a source of drinking water for public or private water supplies.<br />

DSSTox: The Distributed Structure-Searchable Toxicity Database Network, a project of the EPA's<br />

National Center for Computational Toxicology. The DSSTox website provides a public forum for<br />

publishing downloadable, structure-searchable, standardized chemical structure files associated<br />

with chemical inventories or toxicity datasets of environmental relevance. (2)<br />

Effluent: Waste material being discharged into the environment, either treated or untreated. (2)<br />

Environmental justice: The fair treatment of people of all races, cultures, incomes, and<br />

educational levels with respect to the development and enforcement of environmental laws,<br />

255


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

regulations, and policies. The fair distribution of environmental risks across socioeconomic and<br />

racial groups. (2)<br />

Flowback: After the hydraulic fracturing procedure is completed and pressure is released, the<br />

direction of fluid flow reverses, and water and excess proppant flow up through the wellbore to the<br />

surface. The water that returns to the surface is commonly referred to as “flowback.” (3)<br />

Fluid formulation: The entire suite of products and carrier fluid injected into a well during<br />

hydraulic fracturing.<br />

Formation: A geological formation is a body of earth material with distinctive and characteristic<br />

properties and a degree of homogeneity in its physical properties. (2)<br />

Formation water: Water that occurs naturally within the pores of rock. (5)<br />

FracFocus: National registry for chemicals used in hydraulic fracturing, jointly developed by the<br />

Ground Water Protection Council and the Interstate Oil and Gas Compact Commission. Serves as an<br />

online repository where oil and gas well operators can upload information regarding the chemical<br />

compositions of hydraulic fracturing fluids used in specific oil and gas production wells. Also<br />

contains spatial information for well locations and information on well depth and water use.<br />

Geographic information system (GIS): A computer system designed for storing, manipulating,<br />

analyzing, and displaying data in a geographic context, usually as maps. (2)<br />

Gross α: The total radioactivity due to alpha particle emission as inferred from measurements on a<br />

dry sample. (2)<br />

Gross β: The total radioactivity due to beta particle emission as inferred from measurements on a<br />

dry sample. (2)<br />

Ground water: The supply of fresh water found beneath the Earth’s surface, usually in aquifers,<br />

which supply wells and springs. It provides a major source of drinking water. (2)<br />

Halite: A soft, soluble evaporate mineral commonly known as salt or rock salt. Can be critical in<br />

forming hydrocarbon traps and seals because it tends to flow rather than fracture during<br />

deformation, thus preventing hydrocarbons from leaking out of a trap even during and after some<br />

types of deformation. (5)<br />

Hazardous air pollutants: Air pollutants that are not covered by ambient air quality standards but<br />

which, as defined in the Clean Air Act, may present a threat of adverse human health effects or<br />

adverse environmental effects. Although classified as air pollutants, they may also impact drinking<br />

water. (2)<br />

Horizontal drilling: Drilling a portion of a well horizontally to expose more of the formation<br />

surface area to the wellbore. (1)<br />

256


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Hydraulic fracturing: The process of using high pressure to pump sand along with water and<br />

other fluids into subsurface rock formations in order to improve flow of oil and gas into a wellbore.<br />

(1)<br />

Fluid: Specially engineered fluids containing chemical additives and proppant that are<br />

pumped under high pressure into the well to create and hold open fractures in the<br />

formation.<br />

Wastewater: Flowback and produced water, where flowback is the fluid returned to the<br />

surface after hydraulic fracturing has occurred but before the well is placed into production,<br />

and produced water is the fluid returned to the surface after the well has been placed into<br />

production.<br />

Water cycle: The cycle of water in the hydraulic fracturing process, encompassing the<br />

acquisition of water, chemical mixing of the fracturing fluid, injection of the fluid into the<br />

formation, the production and management of flowback and produced water, and the<br />

ultimate treatment and disposal of hydraulic fracturing wastewaters.<br />

Hydraulic gradient: Slope of a water table or potentiometric surface. More specifically, change in<br />

the hydraulic head per unit of distance in the direction of the maximum rate of decrease. (2)<br />

Hydrocarbon: An organic compound containing only hydrogen and carbon, often occurring in<br />

petroleum, natural gas, and coal. (2)<br />

Immiscible: The chemical property where two or more liquids or phases do not readily dissolve in<br />

one another, such as soil and water. (2)<br />

Integrated Risk Information System (IRIS): An electronic database that contains the EPA's latest<br />

descriptive and quantitative regulatory information about chemical constituents. Files on chemicals<br />

maintained in IRIS contain information related to both noncarcinogenic and carcinogenic health<br />

effects. (2)<br />

Laboratory studies: Targeted research conducted to better understand the ultimate fate and<br />

transport of chemical contaminants of concern. The contaminants of concern may be components<br />

of hydraulic fracturing fluids, naturally occurring substances released from the subsurface during<br />

hydraulic fracturing, or treated flowback and produced water that has been released.<br />

Mass spectrometry: Method of chemical analysis in which the substance to be analyzed is heated<br />

and placed in a vacuum. The resulting vapor is exposed to a beam of electrons that causes<br />

ionization to occur, either of the molecules or their fragments. The ionized atoms are separated<br />

according to their mass and can be identified on that basis. (2)<br />

Material Safety Data Sheet (MSDS): Form that contains brief information regarding chemical and<br />

physical hazards, health effects, proper handling, storage, and personal protection appropriate for<br />

use of a particular chemical in an occupational environment. (2)<br />

257


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Monte Carlo simulation: A technique used to estimate the most probable outcomes from a model<br />

with uncertain input data and to estimate the validity of the simulated model.<br />

National Pollution Discharge Elimination System (NPDES): A national program under Section<br />

402 of the Clean Water Act for regulation of discharges of pollutants from point sources to waters of<br />

the United States. Discharges are illegal unless authorized by an NPDES permit. (2)<br />

National Response Center (NRC): Communications center that receives reports of discharges or<br />

releases of hazardous substances into the environment. Run by the US Coast Guard, which relays<br />

information about such releases to the appropriate federal agency. (2)<br />

Natural gas or gas: A naturally occurring mixture of hydrocarbon and non-hydrocarbon gases in<br />

porous formations beneath the Earth’s surface, often in association with petroleum. The principal<br />

constituent of natural gas is methane. (5)<br />

Natural organic matter (NOM): Complex organic compounds that are formed from decomposing<br />

plant animal and microbial material in soil and water. (2)<br />

Offset wells: An existing wellbore close to a proposed well that provides information for planning<br />

the proposed well. (5)<br />

Overburden: Material of any nature, consolidated or unconsolidated, that overlies a deposit of<br />

useful minerals or ores. (2)<br />

Peer review: A documented critical review of a specific major scientific and/or technical work<br />

product. Peer review is intended to uncover any technical problems or unresolved issues in a<br />

preliminary or draft work product through the use of independent experts. This information is then<br />

used to revise the draft so that the final work product will reflect sound technical information and<br />

analyses. The process of peer review enhances the scientific or technical work product so that the<br />

decision or position taken by the EPA, based on that product, has a sound and credible basis.<br />

Permeability: Ability of rock to transmit fluid through pore spaces. (1)<br />

Physicochemical properties: The inherent physical and chemical properties of a molecule such as<br />

boiling point, density, physical state, molecular weight, vapor pressure, etc. These properties define<br />

how a chemical interacts with its environment.<br />

Play: A set of oil or gas accumulations sharing similar geologic, geographic properties, such as<br />

source rock, hydrocarbon type, and migration pathways. (1)<br />

Porosity: Percentage of the rock volume that can be occupied by oil, gas or water. (1)<br />

Produced water: After the drilling and fracturing of the well are completed, water is produced<br />

along with the natural gas. Some of this water is returned fracturing fluid and some is natural<br />

formation water. These produced waters move back through the wellhead with the gas. (4)<br />

258


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Proppant/propping agent: A granular substance (sand grains, aluminum pellets, or other<br />

material) that is carried in suspension by the fracturing fluid and that serves to keep the cracks<br />

open when fracturing fluid is withdrawn after a fracture treatment.<br />

Publicly owned treatment works (POTW): Any device or system used in the treatment (including<br />

recycling and reclamation) of municipal sewage or industrial wastes of a liquid nature that is<br />

owned by a state or municipality. This definition includes sewers, pipes, or other conveyances only<br />

if they convey wastewater to a POTW providing treatment. (2)<br />

Quality assurance (QA): An integrated system of management activities involving planning,<br />

implementation, documentation, assessment, reporting, and quality improvement to ensure that a<br />

process, item, or service is of the type and quality needed and expected by the customer. (2)<br />

Quality assurance project plan (QAPP): A formal document describing in comprehensive detail<br />

the necessary quality assurance procedures, quality control activities, and other technical activities<br />

that need to be implemented to ensure that the results of the work performed will satisfy the stated<br />

performance or acceptance criteria. (2)<br />

Quality Management Plan: A document that describes a quality system in terms of the<br />

organizational structure, policy and procedures, functional responsibilities of management and<br />

staff, lines of authority, and required interfaces for those planning, implementing, documenting, and<br />

assessing all activities conducted. (2)<br />

Radionuclide: Radioactive particle, man-made or natural, with a distinct atomic weight number.<br />

Emits radiation in the form of alpha or beta particles, or as gamma rays. Can have a long life as soil<br />

or water pollutant. Prolonged exposure to radionuclides increases the risk of cancer. (2)<br />

Residuals: The solids generated or retained during the treatment of wastewater. (2)<br />

Safe Drinking Water Act (SDWA): The act designed to protect the nation's drinking water supply<br />

by establishing national drinking water standards (maximum contaminant levels or specific<br />

treatment techniques) and by regulating underground injection control wells. (2)<br />

Scenario evaluation: Exploration of realistic, hypothetical scenarios related to hydraulic fracturing<br />

activities using computer models. Used to identify conditions under which hydraulic fracturing<br />

activities may adversely impact drinking water resources.<br />

Science Advisory Board: A federal advisory committee that provides a balanced, expert<br />

assessment of scientific matters relevant to the EPA. An important function of the Science Advisory<br />

Board is to review EPA’s technical programs and research plans.<br />

Service company: A company that assists well operators by providing specialty services, including<br />

hydraulic fracturing.<br />

Shale: A fine-grained sedimentary rock composed mostly of consolidated clay or mud. Shale is the<br />

most frequently occurring sedimentary rock. (5)<br />

259


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Solubility: The amount of mass of a compound that will dissolve in a unit volume of solution. (2)<br />

Sorption: The act of soaking up or attracting substances. (2)<br />

Source water: Water withdrawn from surface or ground water, or purchased from suppliers, for<br />

hydraulic fracturing.<br />

Spud (spud a well): To start the well drilling process by removing rock, dirt, and<br />

other sedimentary material with the drill bit.<br />

Standard operating procedure (SOP): A written document that details the method of an<br />

operation, analysis, or action whose techniques and procedures are thoroughly prescribed and<br />

which is accepted as the method for performing certain routine or repetitive tasks. (2)<br />

Statistical analysis: Analyzing collected data for the purposes of summarizing information to make<br />

it more usable and/or making generalizations about a population based on a sample drawn from<br />

that population. (2)<br />

Surface water: All water naturally open to the atmosphere (rivers, lakes, reservoirs, ponds,<br />

streams, impoundments, seas, estuaries, etc.). (2)<br />

Surfactant: Used during the hydraulic fracturing process to decrease liquid surface tension and<br />

improve fluid passage through the pipes.<br />

Technical systems audit (TSA): A thorough, systematic, onsite, qualitative audit of facilities,<br />

equipment, personnel, training, procedures, record keeping, data validation, data management, and<br />

reporting aspects of a system. (2)<br />

Tight sands: A geological formation consisting of a matrix of typically impermeable, non-porous<br />

tight sands.<br />

Total dissolved solids (TDS): The quantity of dissolved material in a given volume of water. (2)<br />

Toxicity reference value: A reference point (generally a dose or concentration) where exposures<br />

below that point are not likely to result in an adverse event/effect given a specific range of time.<br />

Toxic Substances Control Act (TSCA): The act that controls the manufacture and sale of certain<br />

chemical substances. (2)<br />

Unconventional resource: An umbrella term for oil and natural gas that is produced by means<br />

that do not meet the criteria for conventional production. What has qualified as unconventional at<br />

any particular time is a complex function of resource characteristics, the available exploration and<br />

production technologies, the economic environment, and the scale, frequency, and duration of<br />

production from the resource. Perceptions of these factors inevitably change over time and often<br />

differ among users of the term. At present, the term is used in reference to oil and gas resources<br />

whose porosity, permeability, fluid trapping mechanism, or other characteristics differ from<br />

conventional sandstone and carbonate reservoirs. Coalbed methane, gas hydrates, shale gas,<br />

fractured reservoirs, and tight gas sands are considered unconventional resources. (5)<br />

260


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

Underground Injection Control (UIC): The program under the Safe Drinking Water Act that<br />

regulates the use of wells to pump fluids into the ground. (2)<br />

Underground injection control well: Units into which hazardous waste is permanently disposed<br />

of by injection 0.25 miles below an aquifer with an underground source of drinking water (as<br />

defined under the SDWA). (2)<br />

Underground source of drinking water: An aquifer currently being used as a source of drinking<br />

water or containing a sufficient quantity of ground water to supply a public water system. USDWs<br />

have a total dissolved solids content of 10,000 milligrams per liter or less and are not aquifers<br />

exempted from protection under the Safe Drinking Water Act. (40 CFR 144.3) (2)<br />

Vapor pressure: The force per unit area exerted by a vapor in an equilibrium state with its pure<br />

solid, liquid, or solution at a given temperature. Vapor pressure is a measure of a substance's<br />

propensity to evaporate. Vapor pressure increases exponentially with an increase in temperature.<br />

(2)<br />

Viscosity: A measure of the internal friction of a fluid that provides resistance to shear within the<br />

fluid. (2)<br />

Volatile: Readily vaporizable at a relatively low temperature. (2)<br />

Wastewater treatment: Chemical, biological, and mechanical procedures applied to an industrial<br />

or municipal discharge or to any other sources of contaminated water in order to remove, reduce,<br />

or neutralize contaminants. (2)<br />

Water withdrawal: The process of taking water from a source and conveying it to a place for a<br />

particular type of use. (2)<br />

Well files: Files that generally contain information regarding all activities conducted at an oil and<br />

gas production well. These files are created by oil and gas operators.<br />

Well operator: A company that ultimately controls and operates oil and gas wells.<br />

261


Study of the Potential Impacts of Hydraulic Fracturing<br />

on Drinking Water Resources: Progress Report December 2012<br />

References<br />

1. Oil and Gas Mineral Services. 2010. Oil and Gas Terminology. Available at<br />

http://www.mineralweb.com/library/oil-and-gas-terms/. Accessed January 20, 2011.<br />

2. US Environmental Protection Agency. 2006. Terminology Services: Terms and Acronyms.<br />

Available at http://iaspub.epa.gov/sor_internet/registry/termreg/home/overview/<br />

home.do. Accessed January 20, 2011.<br />

3. New York State Department of Environmental Conservation. 2011. Supplemental Generic<br />

Environmental Impact Statement on the Oil, Gas and Solution Mining Regulatory Program<br />

(revised draft). Well Permit Issuance for Horizontal Drilling and High-Volume Hydraulic<br />

Fracturing to Develop the Marcellus Shale and Other Low-Permeability Gas Reservoirs.<br />

Available at ftp://ftp.dec.state.ny.us/dmn/download/OGdSGEISFull.pdf. Accessed January<br />

20, 2011.<br />

4. Ground Water Protection Council and ALL Consulting. 2009. Modern Shale Gas<br />

Development in the US: A Primer. Ground Water Protection Council and ALL Consulting for<br />

US Department of Energy. Available at http://www.netl.doe.gov/technologies/oilgas/publications/epreports/shale_gas_primer_2009.pdf.<br />

Accessed December 12, 2012.<br />

5. Schlumberger. Oilfield Glossary. Available at http://www.glossary.oilfield.slb.com/. <br />

Accessed November 11, 2012.<br />

262


this page intentially left blank


this page intentially left blank


EPA 601/R-12/011 | December 2012 | www.epa.gov/<strong>hf</strong>study<br />

PRESORTED STANDARD<br />

POSTAGE & FEES PAID<br />

EPA<br />

PERMIT NO. G-35<br />

Office of Research and Development (8101R)<br />

Washington, DC 20460<br />

Official Business<br />

Penalty for Private Use<br />

$300

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!